100% found this document useful (1 vote)
131 views796 pages

Calculus

Uploaded by

YAAKOV SOLOMON
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
100% found this document useful (1 vote)
131 views796 pages

Calculus

Uploaded by

YAAKOV SOLOMON
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 796

CALCULUS

Gregory Hartman et al.


Virginia Military Institute
Virginia Military Institute
Calculus

Gregory Hartman et al.


This text is disseminated via the Open Education Resource (OER) LibreTexts Project (https://LibreTexts.org) and like the hundreds
of other texts available within this powerful platform, it is freely available for reading, printing and "consuming." Most, but not all,
pages in the library have licenses that may allow individuals to make changes, save, and print this book. Carefully
consult the applicable license(s) before pursuing such effects.
Instructors can adopt existing LibreTexts texts or Remix them to quickly build course-specific resources to meet the needs of their
students. Unlike traditional textbooks, LibreTexts’ web based origins allow powerful integration of advanced features and new
technologies to support learning.

The LibreTexts mission is to unite students, faculty and scholars in a cooperative effort to develop an easy-to-use online platform
for the construction, customization, and dissemination of OER content to reduce the burdens of unreasonable textbook costs to our
students and society. The LibreTexts project is a multi-institutional collaborative venture to develop the next generation of open-
access texts to improve postsecondary education at all levels of higher learning by developing an Open Access Resource
environment. The project currently consists of 14 independently operating and interconnected libraries that are constantly being
optimized by students, faculty, and outside experts to supplant conventional paper-based books. These free textbook alternatives are
organized within a central environment that is both vertically (from advance to basic level) and horizontally (across different fields)
integrated.
The LibreTexts libraries are Powered by MindTouch® and are supported by the Department of Education Open Textbook Pilot
Project, the UC Davis Office of the Provost, the UC Davis Library, the California State University Affordable Learning Solutions
Program, and Merlot. This material is based upon work supported by the National Science Foundation under Grant No. 1246120,
1525057, and 1413739. Unless otherwise noted, LibreTexts content is licensed by CC BY-NC-SA 3.0.
Any opinions, findings, and conclusions or recommendations expressed in this material are those of the author(s) and do not
necessarily reflect the views of the National Science Foundation nor the US Department of Education.
Have questions or comments? For information about adoptions or adaptions contact info@LibreTexts.org. More information on our
activities can be found via Facebook (https://facebook.com/Libretexts), Twitter (https://twitter.com/libretexts), or our blog
(http://Blog.Libretexts.org).

This text was compiled on 04/01/2023


TABLE OF CONTENTS
Licensing

1: Limits
1.1: An Introduction to Limits
1.2: Epsilon-Delta De nition of a Limit
1.3: Finding Limits Analytically
1.4: One Sided Limits
1.5: Continuity
1.6: Limits Involving In nity
1.E: Applications of Limits (Exercises)

2: Derivatives
2.1: Instantaneous Rates of Change- The Derivative
2.2: Interpretations of the Derivative
2.3: Basic Differentiation Rules
2.4: The Product and Quotient Rules
2.5: The Chain Rule
2.6: Implicit Differentiation
2.7: Derivatives of Inverse Functions
2.E: Applications of Derivatives(Exercises)

3: The Graphical Behavior of Functions


3.1: Extreme Values
3.2: The Mean Value Theorem
3.3: Increasing and Decreasing Functions
3.4: Concavity and the Second Derivative
3.5: Curve Sketching
3.E: Applications of the Graphical Behavior of Functions(Exercises)

4: Applications of the Derivative


4.1: Newton's Method
4.2: Related Rates
4.3: Optimization
4.4: Differentials
4.E: Applications of Derivatives (Exercises)

5: Integration
5.1: Antiderivatives and Inde nite Integration
5.2: The De nite Integral
5.3: Riemann Sums
5.4: The Fundamental Theorem of Calculus
5.5: Numerical Integration
5.E: Applications of Integration (Exercises)

1 https://math.libretexts.org/@go/page/43972
6: Techniques of Integration
6.1: Substitution
6.2: Integration by Parts
6.3: Trigonometric Integrals
6.4: Trigonometric Substitution
6.5: Partial Fraction Decomposition
6.6: Hyperbolic Functions
6.7: L'Hopital's Rule
6.8: Improper Integration
6.E: Applications of Antidifferentiation (Exercises)
Index

7: Applications of Integration
7.1: Area Between Curves
7.2: Volume by Cross-Sectional Area- Disk and Washer Methods
7.3: The Shell Method
7.4: Arc Length and Surface Area
7.5: Work
7.6: Fluid Forces
7.E: Applications of Integration (Exercises)

8: Sequences and Series


8.1: Sequences
8.2: In nite Series
8.3: Integral and Comparison Tests
8.4: Ratio and Root Tests
8.5: Alternating Series and Absolute Convergence
8.6: Power Series
8.7: Taylor Polynomials
8.8: Taylor Series
8.E: Applications of Sequences and Series (Exercises)

9: Curves in the Plane


9.1: Conic Sections
9.2: Parametric Equations
9.3: Calculus and Parametric Equations
9.4: Introduction to Polar Coordinates
9.5: Calculus and Polar Functions
9.E: Applications of Curves in a Plane (Exercises)

10: Vectors
10.1: Introduction to Cartesian Coordinates in Space
10.2: An Introduction to Vectors
10.3: The Dot Product
10.4: The Cross Product
10.5: Lines
10.6: Planes
10.E: Applications of Vectors (Exercises)

2 https://math.libretexts.org/@go/page/43972
11: Vector-Valued Functions
11.1: Vector–Valued Functions
11.2: Calculus and Vector-Valued Functions
11.3: The Calculus of Motion
11.4: Unit Tangent and Normal Vectors
11.5: The Arc Length Parameter and Curvature
11.E: Applications of Vector Valued Functions (Exercises)

12: Functions of Several Variables


12.1: Introduction to Multivariable Functions
12.2: Limits and Continuity of Multivariable Functions
12.3: Partial Derivatives
12.4: Differentiability and the Total Differential
12.5: The Multivariable Chain Rule
12.6: Directional Derivatives
12.7: Tangent Lines, Normal Lines, and Tangent Planes
12.8: Extreme Values
12.E: Applications of Functions of Several Variables (Exercises)

13: Multiple Integration


13.1: Iterated Integrals and Area
13.2: Double Integration and Volume
13.3: Double Integration with Polar Coordinates
13.4: Center of Mass
13.5: Surface Area
13.6: Volume Between Surfaces and Triple Integration
13.E: Applications of Multiple Integration (Exercises)

14: Appendix
14.1: Section 1-
14.2: Section 2-
14.3: Section 3-
14.4: Section 4-
14.5: Section 5-
14.6: Section 6-

Index
Glossary
Detailed Licensing

3 https://math.libretexts.org/@go/page/43972
Licensing
A detailed breakdown of this resource's licensing can be found in Back Matter/Detailed Licensing.

1 https://math.libretexts.org/@go/page/115411
CHAPTER OVERVIEW
1: Limits
Calculus means "a method of calculation or reasoning.'' When one computes the sales tax on a purchase, one employs a simple
calculus. When one finds the area of a polygonal shape by breaking it up into a set of triangles, one is using another calculus.
Proving a theorem in geometry employs yet another calculus. Despite the wonderful advances in mathematics that had taken place
into the first half of the 17th century, mathematicians and scientists were keenly aware of what they could not do. (This is true even
today.) In particular, two important concepts eluded mastery by the great thinkers of that time: area and rates of change.
Area seems innocuous enough; areas of circles, rectangles, parallelograms, etc., are standard topics of study for students today
just as they were then. However, the areas of arbitrary shapes could not be computed, even if the boundary of the shape could
be described exactly.
Rates of change were also important. When an object moves at a constant rate of change, then "distance = rate × time.'' But
what if the rate is not constant -- can distance still be computed? Or, if distance is known, can we discover the rate of change?
It turns out that these two concepts were related. Two mathematicians, Sir Isaac Newton and Gottfried Leibniz, are credited with
independently formulating a system of computing that solved the above problems and showed how they were connected. Their
system of reasoning was "a'' calculus. However, as the power and importance of their discovery took hold, it became known to
many as "the'' calculus. Today, we generally shorten this to discuss "calculus.'' The foundation of "the calculus'' is the limit. It is a
tool to describe a particular behavior of a function. This chapter begins our study of the limit by approximating its value graphically
and numerically. After a formal definition of the limit, properties are established that make "finding limits'' tractable. Once the limit
is understood, then the problems of area and rates of change can be approached.
1.1: An Introduction to Limits
1.2: Epsilon-Delta Definition of a Limit
1.3: Finding Limits Analytically
1.4: One Sided Limits
1.5: Continuity
1.6: Limits Involving Infinity
1.E: Applications of Limits (Exercises)

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 1: Limits is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et al. via
source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

1
1.1: An Introduction to Limits
We begin our study of limits by considering examples that demonstrate key concepts that will be explained as we progress.
Consider the function y = sin x

x
. When x is near the value 1, what value (if any) is y near?
While our question is not precisely formed (what constitutes "near the value 1''?), the answer does not seem difficult to find. One
might think first to look at a graph of this function to approximate the appropriate y values.
Consider Figure 1, where y = is graphed. For values of x near 1, it seems that y takes on values near
sin x

x
0.85 . In fact, when
x = 1 , then y = ≈ 0.84 , so it makes sense that when x is "near'' 1, y will be "near'' 0.84.
sin 1

:
FIGURE 1.1 sin(x)/x near x = 1.

Consider this again at a different value for x. When x is near 0, what value (if any) is y near? By considering Figure 1.2, one can
see that it seems that y takes on values near 1. But what happens when x = 0 ? We have
sin 0 0
y → →" ".
0 0

:
FIGURE 1.2 sin(x)/x near x = 0.

The expression "0/0'' has no value; it is indeterminate. Such an expression gives no information about what is going on with the
function nearby. We cannot find out how y behaves near x = 0 for this function simply by letting x = 0 .
Finding a limit entails understanding how a function behaves near a particular value of x. Before continuing, it will be useful to
establish some notation. Let y = f (x); that is, let y be a function of x for some function f . The expression "the limit of y as x
approaches 1'' describes a number, often referred to as L, that y nears as x nears 1. We write all this as

lim y = lim f (x) = L.


x→1 x→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

1.1.1 https://math.libretexts.org/@go/page/4150
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 working 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 approximate limits both graphically and numerically. Graphing a function can provide a good approximation,
though often not very precise. Numerical methods can provide a more accurate approximation. We have already approximated
limits graphically, so we now turn our attention to numerical approximations.
Consider again lim x→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 Figure 1.3.

FIGURE 1.3 : Values of sin(x)/x with x near 1.


Notice that for values of x near 1, we have sin(x)/x near 0.841. The x = 1 row is in bold to highlight the fact that when
considering limits, we are not 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.
Now approximate lim x→0 sin(x)/x numerically. We already approximated the value of this limit as 1 graphically in Figure 1.2.

The table in Figure 1.4 shows the value of sin(x)/x for values of x near 0. Ten places after the decimal point are shown to
highlight how close to 1 the value of sin(x)/x gets as x takes on values very near 0. We include the x = 0 row in bold again to
stress that we are not concerned with the value of our function at x = 0 , only on the behavior of the function near 0.

FIGURE 1.4 : Values of sin(x)/x with x near 1.


This numerical method gives confidence to say that 1 is a good approximation of lim x→0 ; that is,
sin(x)/x

lim sin(x)/x ≈ 1. (1.1.1)


x→0

Later we will be able to prove that the limit is exactly 1.


We now consider several examples that allow us explore different aspects of the limit concept.

1.1.2 https://math.libretexts.org/@go/page/4150
FIGURE 1.5 : Graphically approximating a limit in Example 1.

FIGURE 1.6 : Numerically approximating a limit in Example 1.

Example 1: Approximating the value of a limit

Use graphical and numerical methods to approximate


2
x −x −6
lim . (1.1.2)
2
x→3 6x − 19x + 3

Solution:
To graphically approximate the limit, graph
2 2
y = (x − x − 6)/(6 x − 19x + 3) (1.1.3)

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 Figures 1.5 and 1.6, respectively.
The graph shows that when x is near 3, the value of y is very near 0.3. By considering values of x near 3, we see that
y = 0.294 is a better approximation. The graph and the table imply that

2
x −x −6
lim ≈ 0.294. (1.1.4)
x→3 6 x2 − 19x + 3

This example may bring up a few questions about approximating limits (and the nature of limits themselves).
1. If a graph does not produce as good an approximation as a table, why bother with it?
2. How many values of x in a table are "enough?'' In the previous example, could we have just used x = 3.001 and found a fine
approximation?
Graphs are useful since they give a visual understanding concerning the behavior of a function. Sometimes a function may act
"erratically'' near certain x values which is hard to discern numerically but very plain graphically. Since graphing utilities are very
accessible, it makes sense to make proper use of them.
Since tables and graphs are used only to approximate the value of a limit, there is not a firm answer to how many data points are
"enough.'' Include enough so that a trend is clear, and use values (when possible) both less than and greater than the value in
question. In Example 1, we used both values less than and greater than 3. Had we used just x = 3.001, we might have been
tempted to conclude that the limit had a value of 0.3. While this is not far off, we could do better. Using values "on both sides of 3''
helps us identify trends.

1.1.3 https://math.libretexts.org/@go/page/4150
Example 2: Approximating the value of a limit
Graphically and numerically approximate the limit of f (x) as x approaches 0, where

x +1 x <0
f (x) = { (1.1.5)
2
−x +1 x >0

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 1.7 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.

FIGURE 1.7 : Graphically approximating a limit in Example 2.

FIGURE 1.8 : Numerically approximating a limit in Example 2.


The table shown in Figure 1.8 shows values of f (x) for values of x near 0. It is clear that as x takes on values very near 0,
f (x) takes on values very near 1. It turns out that if we let x = 0 for either "piece'' of f (x), 1 is returned; this is significant and

we'll return to this idea later.


The graph and table allow us to say that lim x→0 f (x) ≈ 1 ; in fact, we are probably very sure it equals 1.

Identifying When Limits Do Not Exist


A function may not have a limit for all values of x. That is, we cannot say lim f (x) = L for some numbers L for all values of
x→c

c , for there may not be a number that f (x) is approaching. There are three ways in which a limit may fail to exist.

1. The function f (x) may approach different values on either side of c .


2. The function may grow without upper or lower bound as x approaches c .
3. The function may oscillate as x approaches c .
We'll explore each of these in turn.

Example 3: Different Values Approached From Left and Right

Explore why lim x→1 f (x) does not exist, where


2
x − 2x + 3 x ≤1
f (x) = { (1.1.6)
x x >1

1.1.4 https://math.libretexts.org/@go/page/4150
FIGURE 1.9 : Observing no limit as x → 1 in Example 3.

FIGURE 1.10 : Values of f (x) near x = 1 in Example 3.


Solution:
A graph of f (x) around x = 1 and a table are given Figures 1.9 and 1.10, respectively. It is clear that as x approaches 1, f (x)
does not seem to approach a single number. Instead, it seems as though f (x) approaches two different numbers. When
considering values of x less than 1 (approaching 1 from the left), it seems that f (x) is approaching 2; when considering values
of x greater than 1 (approaching 1 from the right), it seems that f (x) is approaching 1. Recognizing this behavior is important;
we'll study this in greater depth later. Right now, it suffices to say that the limit does not exist since f (x) is not approaching
one value as x approaches 1.

Example 4: The Function Grows Without Bound

Explore why lim x→1


2
1/(x − 1 ) does not exist.
Solution:
A graph and table of f (x) = 1/(x − 1) are given in Figures 1.11 and 1.12, respectively. Both show that as x approaches 1,
2

f (x) grows larger and larger.

FIGURE 1.11 : Observing no limit as x → 1 in Example 4.

1.1.5 https://math.libretexts.org/@go/page/4150
FIGURE 1.12 : Values of f (x) near x = 1 in Example 4.
We can deduce this on our own, without the aid of the graph and table. If x is near 1, then (x − 1) is very small, and:
2

1
= very large number. (1.1.7)
very small number

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


1
lim (1.1.8)
2
x→1 (x − 1)

does not exist.

Example 5: The Function Oscillates

Explore why lim x→0 sin(1/x) does not exist.


Solution:
Two graphs of f (x) = sin(1/x) are given in Figures 1.13. Figure 1.13(a) shows f (x) on the interval [−1, 1]; notice how f (x)
seems to oscillate near x = 0 . One might think that despite the oscillation, as x approaches 0, f (x) approaches 0. However,
Figure 1.13(b) zooms in on sin(1/x), on the interval [−0.1, 0.1]. Here the oscillation is even more pronounced. Finally, in the
table in Figure 1.13(c), we see sin(x)/x evaluated for values of x near 0. As x approaches 0, f (x) does not appear to approach
any value.

FIGURE 1.13 : Observing that f (x) = sin(1/x) has no limit as x → 0 in Example 5.


It can be shown that in reality, as x approaches 0, sin(1/x) takes on all values between −1 and 1 infinite times! Because of
this oscillation,
limx→0 sin(1/x) does not exist.

Limits of Difference Quotients


We have approximated limits of functions as x approached a particular number. We will consider another important kind of limit
after explaining a few key ideas.

1.1.6 https://math.libretexts.org/@go/page/4150
FIGURE 1.14 : Interpreting a difference quotient as the slope of a secant line.
Let f (x) represent the position function, in feet, of some particle that is 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. Another way of expressing this
is to say
f (1) = 10 and f (5) = 20. (1.1.9)

Since the particle traveled 10 feet in 4 seconds, we can say the particle's average velocity was 2.5 ft/s. We write this calculation
using a "quotient of differences,'' or, a difference quotient:
f (5) − f (1) 10
= = 2.5ft/s. (1.1.10)
5 −1 4

This difference quotient can be thought of as the familiar "rise over run'' used 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 through those two
points. See Figure 1.14.
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 is now

f (1 + h) − f (1) f (1 + h) − f (1)
= . (1.1.11)
(1 + h) − 1 h

Let f (x) = −1.5x + 11.5x; note that f (1) = 10 and f (5) = 20, as in our discussion. We can compute this difference quotient
2

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 1.15. 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 = ? (1.1.12)
h→0 h

FIGURE 1.15 : Secant lines of f (x) at x = 1 and x = 1 + h , for shrinking values of h (i.e., h → 0 ).
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 difference quotient (where the x-axis would represent h values and the y -axis would represent values of
the difference quotient) we settle for making a table. See Figure 1.16. The table gives us reason to assume the value of the limit is
about 8.5.

1.1.7 https://math.libretexts.org/@go/page/4150
FIGURE 1.16 : The difference quotient evaluated at values of h near 0.
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 study of finding limits analytically. In the following
exercises, we continue our introduction and approximate the value of limits.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 1.1: An Introduction to Limits is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory
Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

1.1.8 https://math.libretexts.org/@go/page/4150
1.2: Epsilon-Delta Definition of a Limit
This section introduces the formal definition of a limit. Many refer to this as "the epsilon--delta,'' definition, referring to the letters ϵ
and δ of the Greek alphabet.
Before we give the actual definition, let's consider a few informal ways of describing a limit. Given a function y = f (x) and an x-
value, c , we say that "the limit of the function f , as x approaches c , is a value L'':
1. if "y tends to L'' as "x tends to c .''
2. if "y approaches L'' as "x approaches c .''
3. if "y is near L'' whenever "x is near c .''
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 3. A quick restatement gets us closer to what we want:
3

. If x is within a certain tolerance level of c , then the corresponding value y = f (x) is within a certain tolerance level of L.
The traditional notation for the x-tolerance is the lowercase Greek letter delta, or δ , and the y -tolerance is denoted by lowercase
epsilon, or ϵ. One more rephrasing of 3 nearly gets us to the actual definition:

3
′′
. If x is within δ units of c , then the corresponding value of y is within ϵ units of L.
We can write "x is within δ units of c '' mathematically as
|x − c| < δ, which is equivalent to c − δ < x < c + δ. (1.2.1)

Letting the symbol "⟶'' represent the word "implies,'' we can rewrite 3 as ′′

|x − c| < δ ⟶ |y − L| < ϵ or c − δ < x < c + δ ⟶ L − ϵ < y < L + ϵ. (1.2.2)

The point is that δ and ϵ, being tolerances, can be any positive (but typically small) values. Finally, we have the formal definition of
the limit with the notation seen in the previous section.

Definition 1: The Limit of a Function f

Let I be an open interval containing c , and let f be a function defined on I , except possibly at c . The limit of f (x), as x

approaches c , is L, denoted by
lim f (x) = L, (1.2.3)
x→c

means that given any ϵ > 0 , there exists δ > 0 such that for all x ≠ c , if |x − c| < δ , then |f (x) − L| < ϵ .

(Mathematicians often enjoy writing ideas without using any words. Here is the wordless definition of the limit:
lim f (x) = L ⟺ ∀ ϵ > 0, ∃ δ > 0 s. t. 0 < |x − c| < δ ⟶ |f (x) − L| < ϵ. ) (1.2.4)
x→c

Note the order in which ϵ and δ are given. In the definition, the y -tolerance ϵ is given first and then the limit will exist if we can
find an x-tolerance δ that works.
An example will help us understand this definition. Note that the explanation is long, but it will take one through all steps
necessary to understand the ideas.

Example 6: Evaluating a limit using the definition


Show that lim √−
x = 2.
x→4

Solution:
Before we use the formal definition, let's try some numerical tolerances. What if the y tolerance is 0.5, or ϵ = 0.5 ? How close
to 4 does x have to be so that y is within 0.5 units of 2, i.e., 1.5 < y < 2.5? In this case, we can proceed as follows:

1.2.1 https://math.libretexts.org/@go/page/4151
1.5 < y < 2.5 (1.2.5)

1.5 < √x < 2.5 (1.2.6)

2 2
1.5 < x < 2.5 (1.2.7)

2.25 < x < 6.25. (1.2.8)

So, what is the desired x tolerance? Remember, we want to find a symmetric interval of x values, namely 4 − δ < x < 4 + δ .
The lower bound of 2.25 is 1.75 units from 4; the upper bound of 6.25 is 2.25 units from 4. We need the smaller of these two
distances; we must have δ ≤ 1.75. See Figure 1.17.

FIGURE 1.17 : Illustrating the ϵ − δ process.


Given the y tolerance ϵ = 0.5 , we have found an x tolerance, δ ≤ 1.75, such that whenever x is within δ units of 4, then y is
within ϵ units of 2. That's what we were trying to find.
Let's try another value of ϵ.
What if the y tolerance is 0.01, i.e., ϵ = 0.01 ? How close to 4 does x have to be in order for y to be within 0.01 units of 2 (or
1.99 < y < 2.01)? Again, we just square these values to get 1.99 < x < 2.01 , or
2 2

3.9601 < x < 4.0401. (1.2.9)

What is the desired x tolerance? In this case we must have δ ≤ 0.0399 , which is the minimum distance from 4 of the two
bounds given above.
Note that in some sense, it looks like there are two tolerances (below 4 of 0.0399 units and above 4 of 0.0401 units). However,
we couldn't use the larger value of 0.0401 for δ since then the interval for x would be 3.9599 < x < 4.0401 resulting in y
values of 1.98995 < y < 2.01 (which contains values NOT within 0.01 units of 2).
What we have so far: if ϵ = 0.5 , then δ ≤ 1.75 and if ϵ = 0.01 , then δ ≤ 0.0399. A pattern is not easy to see, so we switch to
general ϵ try to determine δ symbolically. We start by assuming y = √− x is within ϵ units of 2:

1.2.2 https://math.libretexts.org/@go/page/4151
|y − 2| < ϵ

−ϵ < y − 2 < ϵ (Definition of absolute value)


− −
−ϵ < √x − 2 < ϵ (y = √x )

2 − ϵ < √x < 2 + ϵ (Add 2)

2 2
(2 − ϵ) < x < (2 + ϵ) (Square all)

2 2
4 − 4ϵ + ϵ < x < 4 + 4ϵ + ϵ (Expand)

2 2
4 − (4ϵ − ϵ ) < x < 4 + (4ϵ + ϵ ). (Rewrite in the desired form)

The "desired form'' in the last step is "4 − something < x < 4 + something .'' Since we want this last interval to describe an x
tolerance around 4, we have that either δ ≤ 4ϵ − ϵ or δ ≤ 4ϵ + ϵ , whichever is smaller:
2 2

2 2
δ ≤ min{4ϵ − ϵ , 4ϵ + ϵ }. (1.2.10)

Since ϵ > 0 , the minimum is δ ≤ 4ϵ − ϵ . That's the formula: given an ϵ, set δ ≤ 4ϵ − ϵ .


2 2

We can check this for our previous values. If ϵ = 0.5 , the formula gives δ ≤ 4(0.5) − (0.5) 2
= 1.75 and when ϵ = 0.01 , the
formula gives δ ≤ 4(0.01) − (0.01) = 0.399. 2

So given any ϵ > 0 , set δ ≤ 4ϵ − ϵ . Then if |x − 4| < δ (and x ≠ 4 ), then


2
|f (x) − 2| < ϵ , satisfying the definition of the

limit. We have shown formally (and finally!) that lim √x = 2 . x→4

Actually, it is a pain, but this won't work if ϵ ≥ 4 . This shouldn't really occur since ϵ is
supposed to be small, but it could happen. In the cases where ϵ ≥ 4 , just take δ = 1
and you'll be fine.

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.
That is why theorems about limits are so useful! After doing a few more ϵ-δ proofs, you will really appreciate the analytical "short
cuts'' found in the next section.

Example 7: Evaluating a limit using the definition

Show that lim x→2


2
x =4 .
Solution
Let's do this example symbolically from the start. Let ϵ > 0 be given; we want |y − 4| < ϵ , i.e., |x 2
− 4| < ϵ . How do we find
δ such that when |x − 2| < δ , we are guaranteed that | x − 4| < ϵ ?
2

This is a bit trickier than the previous example, but let's start by noticing that |x 2
− 4| = |x − 2| ⋅ |x + 2| . Consider:
ϵ
2
|x − 4| < ϵ ⟶ |x − 2| ⋅ |x + 2| < ϵ ⟶ |x − 2| < . (1.1)
|x + 2|

Could we not set δ = ϵ

|x+2|
?

We are close to an answer, but the catch is that δ must be a constant value (so it can't contain x). There is a way to work around
this, but we do have to make an assumption. Remember that ϵ is supposed to be a small number, which implies that δ will also
be a small value. In particular, we can (probably) assume that δ < 1 . If this is true, then |x − 2| < δ would imply that
|x − 2| < 1 , giving 1 < x < 3 .

Now, back to the fraction ϵ

|x+2|
. If 1 < x < 3 , then 3 < x +2 < 5 (add 2 to all terms in the inequality). Taking reciprocals,
we have

1.2.3 https://math.libretexts.org/@go/page/4151
1 1 1
< < which implies (1.2.11)
5 |x + 2| 3

1 1
< which implies (1.2.12)
5 |x + 2|
ϵ ϵ
< . (1.2)
5 |x + 2|

This suggests that we set δ ≤ ϵ

5
. To see why, let consider what follows 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
2
ϵ
|x − 4| < |x + 2| ⋅ (Combine left side)
5
ϵ ϵ
2
|x − 4| < |x + 2| ⋅ < |x + 2| ⋅ =ϵ (Using (1.2) as long as δ < 1)
5 |x + 2|

We have arrived at |x − 4| < ϵ as desired. Note again, in order to make this happen we needed δ to first be less than 1. That is
2

a safe assumption; we want ϵ to be arbitrarily small, forcing δ to also be small.


We have also picked δ to be smaller than "necessary.'' We could get by with a slightly larger δ , as shown in Figure 1.18. The
dashed outer lines show the boundaries defined by our choice of ϵ. The dotted inner lines show the boundaries defined by
setting δ = ϵ/5 . Note how these dotted lines are within the dashed lines. That is perfectly fine; by choosing x within the dotted
lines we are guaranteed that f (x) will be within ϵ of 4.%If the value we eventually used for δ , namely ϵ/5, is not less than 1,
this proof won't work. For the final fix, we instead set δ to be the minimum of 1 and ϵ/5. This way all calculations above work.

FIGURE 1.18 : Choosing δ = ϵ/5 in Example 7.


In summary, given ϵ > 0 , set δ =≤ ϵ/5 . Then |x − 2| < δ implies |x − 4| < ϵ (i.e. |y − 4| < ϵ ) as desired. This shows that
2

x = 4 . Figure 1.18 gives a visualization of this; by restricting x to values within δ = ϵ/5 of 2, we see that f (x) is
2
lim x→2

within ϵ of 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 assuming
1. |f (x) − L| < ϵ , then perform some algebraic manipulations to give an inequality of the form
2. |x − c| < something.

1.2.4 https://math.libretexts.org/@go/page/4151
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 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.

Example 8: Evaluating a limit using the definition


Prove that lim x 3
− 2x = −1 .
x→1

Solution
We start our scratch--work by considering |f (x) − (−1)| < ϵ :

|f (x) − (−1)| < ϵ (1.2.13)


3
|x − 2x + 1| < ϵ (Now factor) (1.2.14)
2
|(x − 1)(x + x − 1)| <ϵ (1.2.15)
ϵ
|x − 1| < . (1.3)
2
|x + x − 1|

We are at the phase of saying that |x − 1| < something, where something = ϵ/|x 2
+ x − 1| . We want to turn that something
into δ .
Since x is approaching 1, we are safe to assume that x is between 0 and 2. So
0 <x <2
2
0 <x < 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.
2
0 <x +x < 6
2
−1 < x + x − 1 < 5. (subtracted 1 from each part)

In Equation (1.3), we wanted |x − 1| < ϵ/|x 2


+ x − 1| . The above shows that given any x in [0, 2], we know that
2
x +x −1 < 5 which implies that

1 1
< which implies that
2
5 x +x −1
ϵ ϵ
< . (1.4)
2
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 |(x 3
− 2x) − (−1)| < ϵ . We start with |x − 1| < δ :
|x − 1| < δ
ϵ
|x − 1| <
5
ϵ ϵ
|x − 1| < < (for x near 1, from Equation (1.4))
2
5 |x + x − 1|

2
|x − 1| ⋅ | x + x − 1| <ϵ

3
|x − 2x + 1| < ϵ

3
|(x − 2x) − (−1)| < ϵ,

which is what we wanted to show. Thus lim x 3


− 2x = −1 .
x→1

We illustrate evaluating limits once more.

1.2.5 https://math.libretexts.org/@go/page/4151
Example 9: Evaluating a limit using the definition
Prove that lim e x
= 1.
x→0

Solution
Symbolically, we want to take the equation |e x
− 1| < ϵ and unravel it to the form |x − 0| < δ . Here is our scratch--work:
x
|e − 1| < ϵ
x
−ϵ < e −1 < ϵ (Definition of absolute value)
x
1 −ϵ < e < 1 +ϵ (Add 1)

ln(1 − ϵ) < x < ln(1 + ϵ) (Take natural logs)

Making the safe assumption that ϵ < 1 ensures the last inequality is valid (i.e., so that ln(1 − ϵ) is defined). We can then set δ
to be the minimum of | ln(1 − ϵ)| and ln(1 + ϵ) ; i.e.,
δ = min{| ln(1 − ϵ)|, ln(1 + ϵ)} = ln(1 + ϵ). (1.2.16)

Recall ln 1 = 0 and ln x < 0 when 0 < x < 1 . So ln(1 − ϵ) < 0 , hence we consider
its absolute value.
Now, we work through the actual the proof:
|x − 0| < δ

−δ < x < δ (Definition of absolute value)

− ln(1 + ϵ) < x < ln(1 + ϵ).

ln(1 − ϵ) < x < ln(1 + ϵ). (since ln(1 − ϵ) < − ln(1 + ϵ))

The above line is true by our choice of δ and by the fact that since | ln(1 − ϵ)| > ln(1 + ϵ) and ln(1 − ϵ) < 0 , we know
ln(1 − ϵ) < − ln(1 + ϵ) .

x
1 −ϵ < e < 1 +ϵ (Exponentiate)
x
−ϵ < e −1 < ϵ (Subtract 1)

In summary, given ϵ>0 , let δ = ln(1 + ϵ) . Then |x − 0| < δ implies |e


x
− 1| < ϵ as desired. We have shown that
x
lim e = 1.
x→0

We note that we could actually show that lim e =e x→c for any constant c . We do this by factoring out e from both sides,
x c c

leaving us to show lim ex→c = 1 instead. By using the substitution u = x − c , this reduces to showing lim
x−c
e = 1 which u→0
u

we just did in the last example. As an added benefit, this shows that in fact the function f (x) = e is continuous at all values of x,
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
lim(sin x)/x = 1, as we approximated in the previous section.
x→0

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

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 1.2: Epsilon-Delta Definition of a Limit is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by
Gregory Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available
upon request.

1.2.6 https://math.libretexts.org/@go/page/4151
1.3: Finding Limits Analytically
In Section 1.1 we explored the concept of the limit without a strict definition, meaning we could only make approximations. In the
previous section we gave the definition of the limit and demonstrated how to use it to verify our approximations were correct. Thus
far, our method of finding a limit is 1) make a really good approximation either graphically or numerically, and 2) verify our
approximation is correct using a ϵ-δ proof.

This process has its shortcomings, not the least of which is the fact that ϵ--δ proofs are cumbersome. This section gives a series
of theorems which allow us to find limits much more quickly and intuitively.

Recognizing that ϵ-δ proofs are cumbersome, this section gives a series of theorems which allow us to find limits much more
quickly and intuitively.
Suppose that lim f (x) = 2 and lim g(x) = 3 .What is lim(f (x) + g(x)) ? Intuition tells us that the limit should be 5, as we expect
x→2 x→2 x→2

limits to behave in a nice way. The following theorem states that already established limits do behave nicely.

Theorem 1.3.1: Basic Limit Properties

Let b , c , L and K be real numbers, let n be a positive integer, and let f and g be functions with the following limits:

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


x→c x→c

The following limits hold.


1. Constants: lim b = b
x→c

2. Identity: lim x = c
x→c

3. Sums/Differences: lim(f (x) ± g(x)) = L ± K


x→c

4. Scalar Multiples: lim b ⋅ f (x) = bL


x→c

5. Products: lim f (x) ⋅ g(x) = LK


x→c

6. Quotients: lim f (x)/g(x) = L/K , (K ≠ 0)


x→c

7. Powers: lim f (x ) n
=L
n

x→c
−−−−


8. Roots: lim √f (x) = √L
n
n

x→c

9. Compositions: Adjust our previously given limit situation to:


lim f (x) = L, lim g(x) = K and g(L) = K. (1.3.1)
x→c x→L

Then lim g(f (x)) = K. (1.3.2)


x→c

We make a note about Property #8: when n is even, L must be greater than 0. If n is odd, then the statement is true for all L.
We apply the theorem to an example.

Example 1.3.1: Using Basic Limit Properties

Let
2
lim f (x) = 2, lim g(x) = 3 and p(x) = 3 x − 5x + 7. (1.3.3)
x→2 x→2

Find the following limits:


1. lim (f (x) + g(x))
x→2

2. lim (5f (x) + g(x ) 2


)
x→2

3. lim p(x)
x→2

1.3.1 https://math.libretexts.org/@go/page/4152
Solution
1. Using the Sum/Difference rule, we know that lim (f (x) + g(x)) = 2 + 3 = 5 .
x→2

2. Using the Scalar Multiple and Sum/Difference rules, we find that lim (5f (x) + g(x ) 2
) = 5 ⋅ 2 +3
2
= 19.
x→2

3. Here we combine the Power, Scalar Multiple, Sum/Difference and Constant Rules. We show quite a few steps, but in
general these can be omitted:
2
lim p(x) = lim(3 x − 5x + 7)
x→2 x→2

2
= lim 3 x − lim 5x + lim 7
x→2 x→2 x→2

2
=3⋅2 −5 ⋅ 2 +7

=9

Part 3 of the previous example demonstrates how the limit of a quadratic polynomial can be determined using the properties of
Theorem 1. Not only that, recognize that
lim p(x) = 9 = p(2); (1.3.4)
x→2

i.e., the limit at 2 was 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.

Theorem 1.3.2 : Limits of Polynomial and Rational Functions

Limits of Polynomial and Rational Functions}{Let p(x) and q(x) be polynomials and c a real number. Then:
1. lim p(x) = p(c)
x→c
p(x) p(c)
2. lim q(x)
=
q(c)
, where q(c) ≠ 0 .
x→c

Example 1.3.2: Finding a limit of a rational function

Using Theorem 2, find


2
3x − 5x + 1
lim .
4 2
x→−1 x −x +3

Solution
Using Theorem 1.3.2, we can quickly state that
2 2
3x − 5x + 1 3(−1 ) − 5(−1) + 1
lim =
x→−1 4 2 4 2
x −x +3 (−1 ) − (−1 ) +3

9
= = 3.
3

It was likely frustrating in Section 1.2 to do a lot of work to prove that


2
lim x =4 (1.3.5)
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 11.
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 the next section, we will give a formal name to these
functions that behave "nicely.''

1.3.2 https://math.libretexts.org/@go/page/4152
Theorem 1.3.4: Special Limits

Let c be a real number in the domain of the given function and let n be a positive integer. The following limits hold:
x c
1. lim sin x = sin c 4. lim csc x = csc c 7. lim a = a (a > 0) (1.3.6)
x→c x→c x→c

2. lim cos x = cos c 5. lim sec x = sec c 8. lim ln x = ln c (1.3.7)


x→c x→c x→c

n − n
3. lim tan x = tan c 6. lim cot x = cot c 9. lim √x = √c (1.3.8)
x→c x→c x→c

Example 1.3.3: Evaluating limits analytically

Evaluate the following limits.


1. lim cos x
x→π

2. lim(sec 2
x − tan
2
x)
x→3

3. lim cos x sin x


x→π/2

4. lim e ln x

x→1

5. lim sin x

x
x→0

Solution
1. This is a straightforward application of Theorem 3. lim cos x = cos π = −1 .
x→π

2. We can approach this in at least two ways. First, by directly applying Theorem 3, we have:
2 2 2 2
lim(sec x − tan x) = sec 3 − tan 3. (1.3.9)
x→3

Using the Pythagorean Theorem, this last expression is 1; therefore


2 2
lim(sec x − tan x) = 1. (1.3.10)
x→3

We can also use the Pythagorean Theorem from the start.


2 2
lim(sec x − tan x) = lim 1 = 1, (1.3.11)
x→3 x→3

using the Constant limit rule. Either way, we find the limit is 1.
3. Applying the Product limit rule of Theorem 1 and Theorem 3 gives
lim cos x sin x = cos(π/2) sin(π/2) = 0 ⋅ 1 = 0. (1.3.12)
x→π/2

4. Again, we can approach this in two ways. First, we can use the exponential/logarithmic identity that e ln x
=x and evaluate
ln x
lim e = lim x = 1.
x→1 x→1

We can also use the limit Composition Rule of Theorem 1. Using Theorem 3, we have lim ln x = ln 1 = 0 and
x→1

lim e
x
=e
0
=1 , satisfying the conditions of the Composition Rule. Applying this rule,
x→0

ln x x 0
lim e = lim e =e = 1. (1.3.13)
x→1 x→0

Both approaches are valid, giving the same result.


5. We encountered this limit in Section 1.1. Applying our theorems, we attempt to find the limit as
sin x sin 0 " 0 "
lim → → . (1.3.14)
x→0 x 0 0

This, of course, violates a condition of Theorem 1, as the limit of the denominator is not allowed to be 0. Therefore, we are
still unable to evaluate this limit with tools we currently have at hand.

1.3.3 https://math.libretexts.org/@go/page/4152
The section could have been titled "Using Known Limits to Find Unknown Limits.'' By knowing certain limits of functions, we can
find limits involving sums, products, powers, etc., of these functions. We further the development of such comparative tools with
the Squeeze Theorem, a clever and intuitive way to find the value of some limits.
Before stating this theorem formally, suppose we have functions f , g and h where g always takes on values between f and h ; that
is, for all x in an interval,

f (x) ≤ g(x) ≤ h(x). (1.3.15)

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

Theorem 1.3.5: Squeeze Theorem

Let f , g and h be functions on an open interval I containing c such that for all x in I ,
f (x) ≤ g(x) ≤ h(x). (1.3.16)

If
lim f (x) = L = lim h(x), (1.3.17)
x→c x→c

then
lim g(x) = L. (1.3.18)
x→c

It can take some work to figure out appropriate functions by which to "squeeze'' the given function of which you are trying to
evaluate a limit. However, that is generally the only place work is necessary; the theorem makes the "evaluating the limit part'' very
simple.
We use the Squeeze Theorem in the following example to finally prove that lim sin x

x
=1 .
x→0

Example 1.3.4: Using the Squeeze Theorem

Use the Squeeze Theorem to show that


sin x
lim = 1.
x→0 x

Solution
We begin by considering the unit circle. Each point on the unit circle has coordinates (cos θ, sin θ) for some angle θ as shown
in Figure 1.3.1. Using similar triangles, we can extend the line from the origin through the point to the point (1, tan θ), as
shown. (Here we are assuming that 0 ≤ θ ≤ π/2 .Later we will show that we can also consider θ ≤ 0 .)

Figure 1.3.1 : The unit circle and related triangles.


Figure 1.19 shows three regions have been constructed in the first quadrant, two triangles and a sector of a circle, which are
also drawn below. The area of the large triangle is tan θ; the area of the sector is θ/2; the area of the triangle contained inside
1

the sector is sin θ . It is then clear from the diagram that


1

1.3.4 https://math.libretexts.org/@go/page/4152
Multiply all terms by 2

sin θ
, giving
1 θ
≥ ≥ 1. (1.3.19)
cos θ sin θ

Taking reciprocals reverses the inequalities, giving


sin θ
cos θ ≤ ≤ 1. (1.3.20)
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 (1.3.21)
θ→0 θ→0 θ θ→0

sin θ
cos 0 ≤ lim ≤1 (1.3.22)
θ→0 θ

sin θ
1 ≤ lim ≤1 (1.3.23)
θ→0 θ

Clearly this means that lim sin θ

θ
=1
θ→0

Two notes about the previous example 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 maturity, clever proofs like this are easier and easier to create.
Second, this limit tells us more than just that as x approaches 0, sin(x)/x 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 theorem.

Theorem 1.3.5 : Special Limits

1. lim sin x

x
=1
x→0

2. lim cos x−1

x
=0
x→0
1

3. lim(1 + x ) x
=e
x→0
x
e −1
4. lim x
=1
x→0

A short word on how to interpret the latter three limits. We know that as x goes to 0, cos x goes to 1. So, in the second limit, both
the numerator and denominator are approaching 0. However, since the limit is 0, we can interpret this as saying that "cos x is
approaching 1 faster than x is approaching 0.''
In the third limit, inside the parentheses we have an expression that is approaching 1 (though never equaling 1), and we know that 1
raised to any power is still 1. At the same time, the power is growing toward infinity. What happens to a number near 1 raised to a
very large power? In this particular case, the result approaches Euler's number, e , approximately 2.718.
In the fourth limit, we see that as x → 0 , e approaches 1 ``just as fast'' as x → 0 , resulting in a limit of 1.
x

1.3.5 https://math.libretexts.org/@go/page/4152
Our final theorem for this section will be motivated by the following example.

Example 1.3.5: Using algebra to evaluate a limit

Evaluate the following limit:


2
x −1
lim .
x→1 x −1

Solution
We begin by attempting to apply Theorem 3 and substituting 1 for x in the quotient. This gives:
2 2
x −1 1 −1 " 0 "
lim = = , (1.3.24)
x→1 x −1 1 −1 0

and indeterminate form. We cannot apply the theorem.

Figure 1.3.2 : Graphing f in Example 1.3.5 to understand a limit.


By graphing the function, as in Figure 1.3.2, 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:
2
x −1 (x − 1)(x + 1)
= .
x −1 x −1

The function is not defined when x = 1 , but for all other x,

2
(x − 1)(x + 1) (x − 1) (x + 1)
x −1
= = = x + 1. (1.3.25)
x −1 x −1 x −1

Clearly lim x + 1 = 2 . Recall that when considering limits, we are not concerned with the value of the function at 1, only the
x→1

value the function approaches as x approaches 1. Since (x − 1)/(x − 1) and x + 1 are the same at all points except
2
x =1 ,
they both approach the same value as x approaches 1. Therefore we can conclude that
2
x −1
lim = 2.
x→1 x −1

The key to the above example is that the functions y = (x − 1)/(x − 1) and y = x + 1 are identical except at x = 1 . Since limits
2

describe a value the function is approaching, not the value the function actually attains, the limits of the two functions are always
equal.

Theorem 1.3.6 : Limits of Functions Equal At All But One Point

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

lim f (x) = L. (1.3.26)


x→c

1.3.6 https://math.libretexts.org/@go/page/4152
The Fundamental Theorem of Algebra tells us that when dealing with a rational function of the form g(x)/f (x) and directly
g(x)
evaluating the limit lim
f (x)
returns "0/0'', then (x − c) is a factor of both g(x) and f (x). One can then use algebra to factor this
x→c

term out, cancel, then apply Theorem 6. We demonstrate this once more.

Example 1.3.6: Evaluating a limit using Theorem 1.3.6

Evaluate
3 2
x − 2x − 5x + 6
lim .
3 2
x→3 2x + 3x − 32x + 15

Solution
We begin by applying Theorem 3 and substituting 3 for x.This returns the familiar indeterminate form of "0/0''.
Since the numerator and denominator are each polynomials, we know that (x − 3) is factor of each. Using whatever method is
most comfortable to you, factor out (x − 3) from each (using polynomial division, synthetic division, a computer algebra
system, etc.). We find that
3 2 2
x − 2x − 5x + 6 (x − 3)(x + x − 2)
= .
3 2 2
2x + 3x − 32x + 15 (x − 3)(2 x + 9x − 5)

We can cancel the (x − 3) terms as long as x ≠ 3 . Using Theorem 6 we conclude:


3 2 2
x − 2x − 5x + 6 (x − 3)(x + x − 2)
lim = lim
3 2 2
x→3 2x + 3x − 32x + 15 x→3 (x − 3)(2 x + 9x − 5)

2
(x + x − 2)
= lim
2
x→3 (2 x + 9x − 5)

10 1
= = .
40 4

We end this section by revisiting a limit first seen in Section 1.1, a limit of a difference quotient. Let f (x) = −1.5x 2
+ 11.5x; we
f (1+h)−f (1)
approximated the limit lim h
≈ 8.5. We formally evaluate this limit in the following example.
h→0

Example 1.3.7 : Evaluating the limit of a difference quotient


f (1+h)−f (1)
Let f (x) = −1.5x 2
+ 11.5x ; find lim h
.
h→0

Solution
Since f is a polynomial, our first attempt should be to employ Theorem 3 and substitute 0 for h . However, we see that this
gives us "0/0.'' Knowing that we have a rational function hints that some algebra will help. Consider the following steps:
2 2
f (1 + h) − f (1) −1.5(1 + h ) + 11.5(1 + h) − (−1.5(1 ) + 11.5(1))
lim = lim
h→0 h h→0 h
2
−1.5(1 + 2h + h ) + 11.5 + 11.5h − 10
= lim
h→0 h
2
−1.5 h + 8.5h
= lim
h→0 h

h(−1.5h + 8.5)
= lim
h→0 h

= lim(−1.5h + 8.5) (using Theorem 6, as h ≠ 0)


h→0

= 8.5 (using Theorem 3)

This matches our previous approximation.

1.3.7 https://math.libretexts.org/@go/page/4152
This section contains several valuable tools for evaluating limits. One of the main results of this section is Theorem 3; it states that
many functions that we use regularly behave in a very nice, predictable way. In the next section 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.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 1.3: Finding Limits Analytically is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory
Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

1.3.8 https://math.libretexts.org/@go/page/4152
1.4: One Sided Limits
We introduced the concept of a limit gently, approximating their values graphically and numerically. Next came the rigorous
definition of the limit, along with an admittedly tedious method for evaluating them. The previous section gave us tools (which we
call theorems) that allow us to compute limits with greater ease. Chief among the results were the facts that polynomials and
rational, trigonometric, exponential and logarithmic functions (and their sums, products, etc.) all behave "nicely.'' In this section we
rigorously define what we mean by "nicely.''
In Section 1.1 we explored the three ways in which limits of functions failed to exist:
1. The function approached different values from the left and right,
2. The function grows without bound, and
3. The function oscillates.
In this section we explore in depth the concepts behind #1 by introducing the one-sided limit. We begin with formal definitions that
are very similar to the definition of the limit given in Section 1.2, but the notation is slightly different and "x ≠ c '' is replaced with
either "x < c '' or "x > c .''

Definition 2: One Sided Limits


Left-Hand Limit
Let I be an open interval containing c , and let f be a function defined on I , except possibly at c . The limit of f (x) , as x

approaches c from the left, is L, or, the left--hand limit of f at c is L, denoted by


lim f (x) = L, (1.4.1)

x→c

means that given any ϵ > 0 , there exists δ > 0 such that for all x < c , if |x − c| < δ , then |f (x) − L| < ϵ .
Right-Hand Limit
Let I be an open interval containing c , and let f be a function defined on I , except possibly at c . The limit of f (x) , as x

approaches c from the right, is L, or, the right--hand limit of f at c is L, denoted by

lim f (x) = L, (1.4.2)


+
x→c

means that given any ϵ > 0 , there exists δ > 0 such that for all x > c , if |x − c| < δ , then |f (x) − L| < ϵ .

Practically speaking, when evaluating a left-hand limit, we consider only values of x "to the left of c ,'' i.e., where x < c . The
admittedly imperfect notation x → c is used to imply that we look at values of x to the left of c . The notation has nothing to do

with positive or negative values of either x or c . A similar statement holds for evaluating right-hand limits; there we consider only
values of x to the right of c , i.e., x > c . We can use the theorems from previous sections to help us evaluate these limits; we just
restrict our view to one side of c .
We practice evaluating left and right-hand limits through a series of examples.

Example 17: Evaluating one sided limits


x 0 ≤x ≤1
Let f (x) = { , as shown in Figure 1.21. Find each of the following:
3 −x 1 <x <2

1. lim f (x)

x→1

2. lim f (x)
+
x→1

3. lim f (x)
x→1

4. f (1)
5. lim f (x)
+
x→0

6. f (0)
7. lim f (x)

x→2

1.4.1 https://math.libretexts.org/@go/page/4153
8. f (2)

FIGURE 1.21 : A graph of f in Example 17.


Solution
For these problems, the visual aid of the graph is likely more effective in evaluating the limits than using f itself. Therefore we
will refer often to the graph.
1. 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

2. 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 circle'' there; we are evaluating a limit, not the value of the function. Therefore lim f (x) = 2 .
+
x→1

3. The limit of f as x approaches 1 does not exist, as discussed in the first section. The function does not approach one
particular value, but two different values from the left and the right.
4. Using the definition and by looking at the graph we see that f (1) = 1 .
5. As x goes to 0 from the right, we see that f (x) is also approaching 0. Therefore lim f (x) = 0 . Note we cannot consider a
+
x→0

left-hand limit at 0 as f is not defined for values of x < 0 .


6. Using the definition and the graph, f (0) = 0 .
7. As x goes to 2 from the left, we see that f (x) is approaching the value of 1. Therefore lim f (x) = 1.

x→2

8. 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.

Theorem 7: Limits and One Sided Limits


Let f be a function defined on an open interval I containing c . Then
lim f (x) = L (1.4.3)
x→c

if, and only if,


lim f (x) = L and lim f (x) = L. (1.4.4)
− +
x→c x→c

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 17 - 20 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.

Example 18: Evaluating limits of a piecewise-defined function


2 −x 0 <x <1
Let f (x) = { 2
, as shown in Figure 1.22. Evaluate the following.
(x − 2) 1 <x <2

1. lim f (x)

x→1

1.4.2 https://math.libretexts.org/@go/page/4153
2. lim f (x)
+
x→1

3. lim f (x)
x→1

4. f (1)
5. lim f (x)
+
x→0

6. f (0)
7. lim f (x)

x→2

8. f (2)

FIGURE 1.22 : A graph of f from Example 18.


Solution
Again we will evaluate each using both the definition of f and its graph.
1. As x approaches 1 from the left, we see that f (x) approaches 1. Therefore lim f (x) = 1.

x→1

2. As x approaches 1 from the right, we see that again f (x) approaches 1. Therefore lim f (x) = 1 .
x→1+

3. The limit of f as x approaches 1 exists and is 1, as f approaches 1 from both the right and left. Therefore lim f (x) = 1 .
x→1

4. f (1) is not defined. Note that 1 is not in the domain of f as defined by the problem, which is indicated on the graph by an
open circle when x = 1 .
5. As x goes to 0 from the right, f (x) approaches 2. So lim f (x) = 2 .
+
x→0

6. f (0) is not defined as 0 is not in the domain of f .


7. As x goes to 2 from the left, f (x) approaches 0. So lim f (x) = 0

.
x→2

8. f (2) is not defined as 2 is not in the domain of f .

Example 19: Evaluating limits of a piecewise-defined function


2
(x − 1) 0 ≤ x ≤ 2, x ≠ 1
Let f (x) = { , as shown in Figure 1.23. Evaluate the following.
1 x =1

1. lim f (x)

x→1

2. lim f (x)
+
x→1

3. lim f (x)
x→1

4. f (1)

1.4.3 https://math.libretexts.org/@go/page/4153
FIGURE 1.23 : Graphing f in Example 19.
It is clear by looking at the graph that both the left and right-hand limits of f , as x approaches 1, is 0. Thus it is also clear that
the limit is 0; i.e., lim f (x) = 0 . It is also clearly stated that f (1) = 1 .
x→1

Example 20: Evaluating limits of a piecewise-defined function


2
x 0 ≤x ≤1
Let f (x) = { , as shown in Figure 1.24. Evaluate the following.
2 −x 1 <x ≤2

1. lim f (x)

x→1

2. lim f (x)
+
x→1

3. lim f (x)
x→1

4. f (1)

FIGURE 1.24 : Graphing f in Example 20.


Solution
It is clear from the definition of the function and its graph that all of the following are equal:
lim f (x) = lim f (x) = lim f (x) = f (1) = 1. (1.4.5)
− +
x→1 x→1 x→1

In Examples 17 - 20 we were asked to find both lim f (x) and f (1). Consider the following table:
x→1

lim f (x) f (1)


x→1

Example 17 does not exist 1

(1.4.6)
Example 18 1 not defined

Example 19 0 1

Example 20 1 1

Only in Example 20 do both the function and the limit exist and agree. This seems "nice;'' in fact, it seems "normal.'' This is in fact
an important situation which we explore in the next section, entitled "Continuity.'' In short, a continuous function is one in which
when a function approaches a value as x → c (i.e., when lim f (x) = L), it actually attains that value at c . Such functions behave
x→c

nicely as they are very predictable.

1.4.4 https://math.libretexts.org/@go/page/4153
Contributors and Attributions
Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 1.4: One Sided Limits is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et
al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

1.4.5 https://math.libretexts.org/@go/page/4153
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 (1.5.1)
x→1

then as x is close to 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.

Definition 3 Continuous Function


Let f be a function defined on an open interval I containing c .
1. f is continuous at c if lim f (x) = f (c).
x→c

2. f is continuous on I if f is continuous at c for all values of c in I . If f is continuous on (−∞, ∞), we say f is


continuous everywhere.

A useful way to establish whether or not a function f is continuous at c is to verify the following three things:
1. lim f (x) exists,
x→c

2. f (c) is defined, and


3. lim f (x) = f (c).
x→c

Example 21: Finding intervals of continuity


Let f be defined as shown in Figure 1.25. Give the interval(s) on which f is continuous.

FIGURE 1.25 : A graph of f in Example 21.


Solution
We proceed by examining the three criteria for continuity.
1. The limits lim f (x) exists for all c between 0 and 3.
x→c

2. f (c) is defined for all c between 0 and 3, except for c = 1 . We know immediately that f cannot be continuous at x = 1 .
3. The limit lim f (x) = f (c) for all c between 0 and 3, except, of course, for c = 1 .
x→c

We conclude that f is continuous at every point of (0, 3) except at x = 1 . Therefore f is continuous on (0, 1) ∪ (1, 3).

Example 22: Finding intervals of continuity


The floor function, f (x) = ⌊x⌋, returns the largest integer smaller than the input x . (For example, f (π) = ⌊π⌋ = 3 .) The
graph of f in Figure 1.26 demonstrates why this is often called a "step function.''
Give the intervals on which f is continuous.

1.5.1 https://math.libretexts.org/@go/page/4154
FIGURE 1.26 : A graph of the step function in Example 22.
Solution
We examine the three criteria for continuity.
1. The limits lim f (x) do not exist at the jumps from one "step'' to the next, which occur at all integer values of c . Therefore
x→c

the limits exist for all c except when c is an integer.


2. The function is defined for all values of c .
3. The limit lim f (x) = f (c) for all values of c where the limit exist, since each step consists of just a line.
x→c

We conclude that f is continuous everywhere except at integer values of c . So the intervals on which f is continuous are
… , (−2, −1), (−1, 0), (0, 1), (1, 2), … . (1.5.2)

Our definition of continuity on an interval specifies the interval is an open interval. We can extend the definition of continuity to
closed intervals by considering the appropriate one-sided limits at the endpoints.

Definition 4: Continuity on Closed Intervals


Let f be defined on the closed interval [a, b] for some real numbers a, b. f is continuous on [a, b] if:
1. f is continuous on (a, b),
2. lim f (x) = f (a) and
+
x→a

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



x→b

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

Example 23: Determining intervals on which a function is continuous


For each of the following functions, give the domain of the function and the interval(s) on which it is continuous.
1. f (x) = 1/x
2. f (x) = sin x
3. f (x) = √−x
− −−−−
4. f (x) = √1 − x 2

5. f (x) = |x|
Solution
We examine each in turn.
1. The domain of f (x) = 1/x is (−∞, 0) ∪ (0, ∞) . As it is a rational function, we apply Theorem 2 to recognize that f is
continuous on all of its domain.
2. The domain of f (x) = sin x is all real numbers, or (−∞, ∞). Applying Theorem 3 shows that sin x is continuous
everywhere.
3. The domain of f (x) = √− −
x is [0, ∞). Applying Theorem 3 shows that f (x) = √x is continuous on its domain of [0, ∞).
−−−− −
4. The domain of f (x) = √1 − x is [−1, 1]. Applying Theorems 1 and 3 shows that f is continuous on all of its domain,
2

[−1, 1].

1.5.2 https://math.libretexts.org/@go/page/4154
−x x <0
5. The domain of f (x) = |x| is (−∞, ∞). We can define the absolute value function as f (x) = { . Each "piece''
x x ≥0

of this piecewise defined function is continuous on all of its domain, giving that f is continuous on (−∞, 0) and [0, ∞).
We cannot assume this implies that f is continuous on (−∞, ∞); we need to check that lim f (x) = f (0), as x = 0 is the
x→0

point where f transitions from one ``piece'' of its definition to the other. It is easy to verify that this is indeed true, hence we
conclude that f (x) = |x| is continuous everywhere.

Continuity is inherently tied to the properties of limits. Because of this, the properties of limits found in Theorems 1 and 2 apply to
continuity as well. Further, now knowing the definition of continuity we can re--read Theorem 3 as giving a list of functions that
are continuous on their domains. The following theorem states how continuous functions can be combined to form other
continuous functions, followed by a theorem which formally lists functions that we know are continuous on their domains.

Theorem 8: Properties of Continuous Functions


Let f and g be continuous functions on an interval I , let c be a real number and let n be a positive integer. The following
functions are continuous on I .
1. Sums/Differences : f ± g
2. Constant Multiples : c ⋅ f
3. Products : f ⋅ g
4. Quotients : f /g (as long as g ≠ 0 on I )
5. Powers : f n


6. Roots : √f (if n is even then f ≥ 0 on I ; if n is odd, then true for all values of f on I .)
n

7. Compositions : Adjust the definitions of f and g to: Let f be continuous on I , where the range of f on I is J , and let g be
continuous on J . Then g ∘ f , i.e., g(f (x)), is continuous on I .

Theorem 9: Continuous Functions


The following functions are continuous on their domains.
1. f (x) = sin x 2. f (x) = cos x (1.5.3)

3. f (x) = tan x 4. f (x) = cot x (1.5.4)

5. f (x) = sec x 6. f (x) = csc x (1.5.5)

n

7. f (x) = ln x 8. f (x) = √x , (1.5.6)
x
9. f (x) = a (a > 0) (where n is a positive integer) (1.5.7)

We apply these theorems in the following Example.

Example 24: Determining intervals on which a function is continuous


State the interval(s) on which each of the following functions is continuous.
−− −−− − −−−−
1. f (x) = √ x − 1 + √ 5 − x 3. f (x) = tan x (1.5.8)
−−−
2. f (x) = x sin x 4. f (x) = √ln x (1.5.9)

FIGURE 1.27 : A graph of f in Example 24(a).


Solution

1.5.3 https://math.libretexts.org/@go/page/4154
We examine each in turn, applying Theorems 8 and 9 as appropriate.
1. The square--root terms are continuous on the intervals [1, ∞) and (−∞, 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
1.27.
2. The functions y = x and y = sin x are each continuous everywhere, hence their product is, too.
3. Theorem 9 states that f (x) = tan x is continuous "on its domain.'' Its domain includes all real numbers except odd
multiples of π/2. Thus f (x) = tan x is continuous on
3π π π π π 3π
… (− ,− ), (− , ), ( , ),…, (1.5.10)
2 2 2 2 2 2

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


π

4. The domain of y = √− x is [0, ∞). The range of y = ln x is (−∞, ∞) , but if we restrict its domain to [1, ∞) its range is

[0, ∞). So restricting y = ln x to the domain of [1, ∞) restricts its output is [0, ∞), on which y = √x is defined. Thus the
−−−
domain of f (x) = √ln x is [1, ∞).

A common way of thinking of a continuous function is that "its graph can be sketched without lifting your pencil.'' That is, its
graph forms a "continuous'' curve, without holes, breaks or jumps. While beyond the scope of this text, this pseudo--definition
glosses over some of the finer points of continuity. Very strange functions are continuous that one would be hard pressed to actually
sketch by hand.
This intuitive notion of continuity does help us understand another important concept as follows. Suppose f is defined on [1, 2] and
f (1) = −10 and f (2) = 5 . If f is continuous on [1, 2] (i.e., its graph can be sketched as a continuous curve from (1, −10) to

(2, 5)) then we know intuitively that somewhere on [1, 2] f must be equal to −9, and −8, and −7, − 6, … , 0, 1/2, etc. In

short, f takes on all intermediate values between −10 and 5. It may take on more values; f may actually equal 6 at some time, for
instance, but we are guaranteed all values between −10 and 5.
While this notion seems intuitive, it is not trivial to prove and its importance is profound. Therefore the concept is stated in the
form of a theorem.

Theorem 10: Intermediate Value Theorem


Let f be a continuous function on [a, b] and, without loss of generality, let f (a) < f (b) . Then for every value y , where
f (a) < y < f (b) , there is a value c in [a, b] such that f (c) = y .

One important application of the Intermediate Value Theorem is root finding. 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 approximations 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 a c in [a, b] such that f (c) = 0 . The theorem does not give us any clue

as to where that value is in the interval [a, b], just that it exists.
There is a technique that produces a good approximation of c . Let d be the midpoint of the interval [a, b] 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.
2. f (d) < 0 Then we know there is a root of f on the interval [d, b] -- we have halved the size of our interval, hence are closer to a
good approximation of the root.
3. f (d) > 0 Then we know there is a root of f on the interval [a, d] -- again,we have halved the size of our interval, hence are
closer to a good approximation of the root.
Successively applying this technique is called the Bisection Method of root finding. We continue until the interval is sufficiently
small. We demonstrate this in the following example.

Example 25: Using the Bisection Method

Approximate the root of f (x) = x − cos x , accurate to three places after the decimal.

1.5.4 https://math.libretexts.org/@go/page/4154
FIGURE 1.28 : Graphing a root of f (x) = x − cos x .
Solution
Consider the graph of f (x) = x − cos x , shown in Figure 1.28. 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.


1. Iteration 1: f (0.7) < 0, f (0.9) > 0, and f (0.8) > 0. So replace 0.9 with 0.8 and repeat.
2. Iteration 2: f (0.7) < 0, f (0.8) > 0, and at the midpoint, 0.75, we have f (0.75) > 0. So replace 0.8 with 0.75 and repeat.
Note that we don't need to continue to check the endpoints, just the midpoint. Thus we put the rest of the iterations in Table
1.29.

FIGURE 1.29 : Iterations of the Bisection Method of Root Finding


Notice that in the 12 iteration we have the endpoints of the interval each starting with 0.739. Thus we have narrowed the
th

zero down to an accuracy of the first three places after the decimal. Using a computer, we have
f (0.7390) = −0.00014, f (0.7391) = 0.000024. (1.5.11)

Either endpoint of the interval gives a good approximation of where f is 0. The Intermediate Value Theorem states that the
actual zero is still within this 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 endpoints differ by a preset small amount. One of the authors did write such a program and found the zero of f , accurate to
10 places after the decimal, to be 0.7390851332. While it took a few minutes to write the program, it took less than a
thousandth of a second for the program to run the necessary 35 iterations. In less than 8 hundredths of a second, the zero was
calculated to 100 decimal places (with less than 200 iterations).

It is a simple matter to extend the Bisection Method to solve problems similar to "Find x, where f (x) = 0 .'' For instance, we can
find x, where f (x) = 1 . It actually works very well to define a new function g where g(x) = f (x) − 1 . Then use the Bisection
Method to solve g(x) = 0 .
Similarly, given two functions f and g , we can use the Bisection Method to solve f (x) = g(x) . Once again, create a new function
h where h(x) = f (x) − g(x) and solve h(x) = 0 .

In Section 4.1 another equation solving method will be introduced, called Newton's Method. In many cases, Newton's Method is
much faster. It relies on more advanced mathematics, though, so we will wait before introducing it.

1.5.5 https://math.libretexts.org/@go/page/4154
This section formally defined what it means to be a continuous function. "Most'' functions that we deal with are continuous, so
often it feels odd to have to formally define this concept. Regardless, it is important, and forms the basis of the next chapter.
In the next section we examine one more aspect of limits: limits that involve infinity.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 1.5: Continuity is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et al. via
source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

1.5.6 https://math.libretexts.org/@go/page/4154
1.6: Limits Involving Infinity
In Definition 1 we stated that in the equation lim f (x) = L , both c and L were numbers. In this section we relax that definition a
x→c

bit by considering situations when it makes sense to let c and/or L be "infinity.''


As a motivating example, consider f (x) = 1/x , as shown in Figure 1.30. Note how, as
2
x approaches 0, f (x) grows very, very
large. It seems appropriate, and descriptive, to state that
1
lim = ∞. (1.6.1)
x→0 x2

Also note that as x gets very large, f (x) gets very, very small. We could represent this concept with notation such as
1
lim = 0. (1.6.2)
x→∞ x2

FIGURE 1.30 : Graphing f (x) = 1/x for values of x near 0 .


2

We explore both types of use of ∞ in turn.

Definition 5: Limit of infinity


We say lim f (x) = ∞ if for every M >0 there exists δ > 0 such that for all x ≠ c , if |x − c| < δ , then f (x) ≥ M .
x→c

This is just like the ϵ--δ definition 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 f (c). Here, given any (large) value M , if we let x get close enough
to c (within δ units of c ), then f (x) will be at least as large as M . In other words, if we get close enough to c , then we can make
f (x) as large as we want. We can define limits equal to −∞ in a similar way.

It is important to note that by saying lim f (x) = ∞ we are implicitly stating that \textit{the} limit of f (x), as x approaches c , does
x→c

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.

Example 26: Evaluating limits involving infinity


Find lim 1
2
as shown in Figure 1.31.
x→1 (x−1)

FIGURE 1.31 : Observing infinite limit as x → 1 in Example 26.

1.6.1 https://math.libretexts.org/@go/page/4155
Solution
In Example 4 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 (.99) = 10 , f (.999) = 10 , f (.9999) = 10 . A similar thing happens on the other side of 1. In general, let a
4 6 8

−− −−
"large'' value M be given. Let δ = 1/√M . If x is within δ of 1, i.e., if |x − 1| < 1/√M , then:
1
|x − 1| <
−−
√M

1
2
(x − 1) <
M
1
> M,
2
(x − 1)

which is what we wanted to show. So we may say lim 1/(x − 1) 2


=∞ .
x→1

Example 27: Evaluating limits involving infinity


Find lim , as shown in Figure 1.32.
1

x
x→0

FIGURE 1.32 : Evaluating lim . 1

x
x→0

Solution
It is easy to see that the function grows without bound near 0, but it does so in different ways on different sides of 0. Since its
behavior is not consistent, we cannot say that lim = ∞ . However, we can make a statement about one--sided limits. We can
1

x
x→0

state that lim


+
1

x
=∞ and lim

1

x
= −∞ .
x→0 x→0

Vertical Asymptotes
If the limit of f (x) as x approaches c from either the left or right (or both) is ∞ or −∞ , we say the function has a vertical
asymptote at c .

Example 28: Finding vertical asymptotes


3x
Find the vertical asymptotes of f (x) = 2
.
x −4

1.6.2 https://math.libretexts.org/@go/page/4155
FIGURE 1.33 : Graphing f (x) = 3x

x2 −4
.

Solution
Vertical asymptotes occur where the function grows without 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 1.33. Thus the vertical
asymptotes are at x = ±2 .

When a rational function has a vertical asymptote at x = c , we can conclude that the denominator is 0 at x = c . However, just
because the denominator is 0 at a certain point does not mean there is a vertical asymptote there. For instance,
f (x) = (x − 1)/(x − 1) does not have a vertical asymptote at x = 1 , as shown in Figure 1.34. While the denominator does get
2

small near x = 1 , the numerator gets small too, matching the denominator step for step. In fact, factoring the numerator, we get
(x − 1)(x + 1)
f (x) = . (1.6.3)
x −1

Canceling the common term, we get that f (x) = x + 1 for x ≠ 1 . So there is clearly no asymptote, rather a hole exists in the graph
at x = 1 .

FIGURE 1.34 : Graphically showing that f (x) = x −1

x−1
does not have an asymptote at x = 1 .
The above example may seem a little contrived. Another example demonstrating this important concept is f (x) = (sin x)/x. We
have considered this function several times in the previous sections. We found that lim = 1 ; i.e., there is no vertical
sin x

x
x→0

asymptote. No simple algebraic cancellation makes this fact obvious; we used the Squeeze Theorem in Section 1.3 to prove this.
If the denominator is 0 at a certain point but the numerator is not, then there will usually be a vertical asymptote at that point. On
the other hand, if the numerator and denominator are both zero at that point, then there may or may not be a vertical asymptote at
that point. This case where the numerator and denominator are both zero returns us to an important topic.

Indeterminate Forms
We have seen how the limits
2
sin x x −1
lim and lim (1.6.4)
x→0 x x→1 x −1

1.6.3 https://math.libretexts.org/@go/page/4155
each return the indeterminate form "0/0'' when we blindly plug in x = 0 and x = 1 , respectively. However, 0/0 is not a valid
arithmetical expression. It gives no indication that the respective limits are 1 and 2.
With a little cleverness, one can come up 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 actual 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 canceling) or it may require a tool such as the Squeeze Theorem. In a later section we will learn a technique called
l'Hospital's Rule that provides another way to handle indeterminate forms.
Some other common indeterminate forms are ∞ − ∞ , ∞ ⋅ 0 , ∞/∞, 0 , ∞ and 1 . Again, keep in mind that these are the
0 0 ∞

"blind'' results of evaluating a limit, 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 without 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.

Limits at Infinity and Horizontal Asymptotes


At the beginning of this section we briefly considered what happens to f (x) = 1/x as x grew very large. Graphically, it concerns
2

the behavior of the function to the "far right'' of the graph. We make this notion more explicit in the following definition.

Definition 6: Limits at Infinity and Horizontal Asymptote


1. We say lim f (x) = L if for every ϵ > 0 there exists M >0 such that if x ≥ M , then |f (x) − L| < ϵ .
x→∞

2. We say lim f (x) = L if for every ϵ > 0 there exists M <0 such that if x ≤ M , then |f (x) − L| < ϵ .
x→−∞

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


x→∞ x→−∞

We can also define limits such as lim f (x) = ∞ by combining this definition with Definition 5.
x→∞

Example 29: Approximating horizontal asymptotes


2

Approximate the horizontal asymptote(s) of f (x) = x2 +4


x
.

Solution
We will approximate the horizontal asymptotes by approximating the limits
2 2
x x
lim and lim . (1.6.5)
x→−∞ 2 x→∞ 2
x +4 x +4

Figure 1.35(a) shows a sketch of f , and part (b) gives values of f (x) for large magnitude values of x. It seems reasonable to
conclude from both of these sources that f has a horizontal asymptote at y = 1 .

1.6.4 https://math.libretexts.org/@go/page/4155
FIGURE 1.35 : Using a graph and a table to approximate a horizontal asymptote in Example 29.
Later, we will show how to determine this analytically.

Horizontal asymptotes can take on a variety of forms. Figure 1.36(a) shows that f (x) = x/(x 2
+ 1) has a horizontal asymptote of
y = 0 , where 0 is approached from both above and below.

−−−−−
Figure 1.36(b) shows that f (x) = x/√x 2
+1 has two horizontal asymptotes; one at y = 1 and the other at y = −1 .
Figure 1.36(c) shows that f (x) = (sin x)/x has even more interesting behavior than at just x =0 ; as x approaches ±∞ , f (x)

approaches 0, but oscillates as it does this.

FIGURE 1.36 : Considering different types of horizontal asymptotes.


We can analytically evaluate limits at infinity for rational functions once we understand lim 1/x . As x gets larger and larger, the
x→∞

1/x gets smaller and smaller, approaching 0. We can, in fact, make 1/x as small as we want by choosing a large enough value of x.
Given ϵ, we can make 1/x < ϵ by choosing x > 1/ϵ. Thus we have lim 1/x = 0.
x→∞

It is now not much of a jump to conclude the following:


1 1
lim =0 and lim =0 (1.6.6)
n n
x→∞ x x→−∞ x

Now suppose we need to compute the following limit:


3
x + 2x + 1
lim . (1.6.7)
x→∞ 3 2
4x − 2x +9

A good way of approaching this is to divide through the numerator and denominator by x
3
(hence dividing by 1), which is the
largest power of x to appear in the function. Doing this, we get
3 3 3
x + 2x + 1 1/x x + 2x + 1
lim = lim ⋅
x→∞ 3 2 x→∞ 3 3 2
4x − 2x +9 1/x 4x − 2x +9
3 3 3 3
x /x + 2x/ x + 1/ x
= lim
x→∞ 4 x3 / x3 − 2 x2 / x3 + 9/ x3

2 3
1 + 2/ x + 1/ x
= 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/x , we see that the limit n

becomes

1.6.5 https://math.libretexts.org/@go/page/4155
1 +0 +0 1
= . (1.6.8)
4 −0 +0 4

This procedure works for any rational function. In fact, it gives us the following theorem.

Theorem 11: Limits of Rational Functions at Infinity


Let f (x) be a rational function of the following form:
n n−1
an x + an−1 x + ⋯ + a1 x + a0
f (x) = , (1.6.9)
m m−1
bm x + bm−1 x + ⋯ + b1 x + b0

where any of the coefficients may be 0 except for a and b . n m

an
1. If n = m , then lim f (x) = lim f (x) =
bm
.
x→∞ x→−∞

2. If n < m , then lim f (x) = lim f (x) = 0 .


x→∞ x→−∞

3. If n > m , then lim f (x) and lim f (x) are both infinite.
x→∞ x→−∞

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 x and in the limit all the terms will approach 0 except for a x /x
n
n
n n

and b x /x . Since n = m , this will leave us with the limit a /b . If n < m , then after dividing through by x , all the terms in
m
m n
n m
m

the numerator will approach 0 in the limit, leaving us with 0/b or 0. If n > m , and we try dividing through by x , we end up
m
n

with all the terms in the denominator tending toward 0, while the x term in the numerator does not approach 0. This is indicative
n

of some sort of infinite limit.


Intuitively, as x gets very large, all the terms in the numerator are small in comparison to a x , and likewise all the terms in the n
n

denominator are small compared to b x . If n = m , looking only at these two important terms, we have (a x )/(b x ). This
n
m
n
n
n
m

reduces to a /b . If n < m , the function behaves like a /(b x


n m ), which tends toward 0. If n > m , the function behaves like
n m
m−n

/ b , which will tend to either ∞ or −∞ depending on the values of n , m , a , b and whether you are looking for
n−m
a x
n m n m

lim f (x) or lim f (x).


x→∞ x→−∞

With care, we can quickly evaluate limits at infinity for a large number of functions by considering the largest powers of x. For
instance, consider again lim , graphed in Figure ??? (b). When x is very large, x + 1 ≈ x . Thus
x 2 2

x→±∞ √x2 +1

− −−−− −− x x
2 2
√ x + 1 ≈ √x = |x|, and ≈ . (1.6.10)
− −−−−
√ x2 + 1 |x|

This expression is 1 when x is positive and −1 when x is negative. Hence we get asymptotes of y = 1 and y = −1 , respectively.

Example 30: Finding a limit of a rational function


2

Confirm analytically that y = 1 is the horizontal asymptote of f (x) = x +4


x
2
, as approximated in Example 29.
Solution
Before using Theorem 11, let's use the technique of evaluating limits at infinity of rational functions that led to that theorem.
The largest power of x in f is 2, so divide the numerator and denominator of f by x , then take limits. 2

2 2 2
x x /x
lim = lim
x→∞ 2 x→∞ 2 2 2
x +4 x /x + 4/ x

1
= lim
x→∞ 2
1 + 4/x

1
=
1 +0

= 1.

We can also use Theorem 11 directly; in this case n =m so the limit is the ratio of the leading coefficients of the numerator
and denominator, i.e., 1/1 = 1.

1.6.6 https://math.libretexts.org/@go/page/4155
Example 31: Finding limits of rational functions
Use Theorem 11 to evaluate each of the following limits.
2 2
x + 2x − 1 x −1
1. lim 3. lim (1.6.11)
3
x→−∞ x +1 x→∞ 3 −x
2
x + 2x − 1
2. lim (1.6.12)
x→∞ 1 − x − 3x2

FIGURE 1.37 : Visualizing the functions in Example 31.


Solution
1. The highest power of x is in the denominator. Therefore, the limit is 0; see Figure 1.37(a).
2. The highest power of x is x , which occurs in both the numerator and denominator. The limit is therefore the ratio of the
2

coefficients of x , which is −1/3. See Figure 1.37(b).


2

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 x and −x. The expression in the limit will behave like x /(−x) = −x
2 2

for large values of x. Therefore, the limit is −∞ . See Figure 1.37(c).

Chapter Summary
In this chapter we:
defined the limit,
found accessible ways to approximate their values numerically and graphically,
developed a not--so--easy method of proving the value of a limit (ϵ − δ proofs),
explored when limits do not exist,
defined continuity and explored properties of continuous functions, and
considered limits that involved infinity.
Why? Mathematics is famous for building on itself and calculus proves to be no exception. In the next chapter we will be interested
in "dividing by 0.'' That is, we will want to divide a quantity by a smaller and smaller number and see what value the quotient
approaches. In other words, we will want to find a limit. These limits will enable us to, among other things, determine exactly how
fast something is moving when we are only given position information.
Later, we will want to add up an infinite list of numbers. We will do so by first adding up a finite list of numbers, then take a limit
as the number of things we are adding approaches infinity. Surprisingly, this sum often is finite; that is, we can add up an infinite
list of numbers and get, for instance, 42.
These are just two quick examples of why we are interested in limits. Many students dislike this topic when they are first
introduced to it, but over time an appreciation is often formed based on the scope of its applicability.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 1.6: Limits Involving Infinity is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory
Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

1.6.7 https://math.libretexts.org/@go/page/4155
1.E: Applications of Limits (Exercises)
1.1: An Introduction to Limits
Terms and Concepts
1. In your own words, what does it mean to "find the limit of f (x) as x approaches 3"?
2. An expression of the form is called _____. 0

3. T/F: 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

5. In your own words, what is a difference quotient.

Problems
In Exercises 6-16, approximate the given limits both numerically and graphically.
6. lim x 2
+ 3x − 5
x→1

7. lim x 3
− 3x
2
+x −5
x→0

8. lim x+1
2
x +3x
x→0

9. lim x −2x−3
2
x −4x+3
x→3

2
x +8x+7
10. lim 2
x +6x+5
x→−1

11. lim x +7x+10

x2 −4x+4
x→2

x +2 x ≤2
12. lim, where f (x) = { .
x→2 3x − 5 x >2

2
x −x +1 x ≤3
13. lim, where f (x) = { .
x→3 2x + 1 x >3

cos x x ≤0
14. lim, where f (x) = { 2
.
x→0 x + 3x + 1 x >0

sin x x ≤ π/2
15. lim , where f (x) = { .
x→π/2 cos x x > π/2

f (a+h)−f (a)
In Exercises 16-24, a function f and a value a are given. Approximate the limit of the difference quotient, lim
h
,
h→0

using h = ±0.1, ±0.01.

16. f (x) = −7x + 2, a =3

17. f (x) = 9x + 0.06, a = −1

18. f (x) = x 2
+ 3x − 7, a =1

19. f (x) = 1

x+1
, a =2

20. f (x) = −4x 2


+ 5x − 1, a = −3

21. f (x) = ln x, a =5

22. f (x) = sin x, a =π

23. f (x) = cos x, a =π

1.2: Epsilon-Delta Definition of a Limit

1.E.1 https://math.libretexts.org/@go/page/9966
Terms and Concepts
1. What is wrong with the following "definition" of a limit?
"The limit of f (x), as x approaches a , is K means that given any
′′
δ >0 there exists ϵ>0 such that whenever |f (x) − K| < ϵ ,
we have |x − a| < δ .
2. Which is given first in establishing a limit, the x-tolerance or the y-tolerance?
3. T/F: ϵ must always be positive.
4. T/F: δ must always be positive.

Problems
In Exercises 5-11, prove the given limit using an ϵ − δ proof.
5. lim 3 − x + −2
x→5

6. lim x 2
−3 = 6
x→3

7. lim x 2
+ x − 5 = 15
x→4

8. lim x 3
−1 = 7
x→2

9. lim 5 = 5
x→2

10. lim e 2x
−1 = 0
x→0

11. lim sin x = 0 (Hint: use the fact that | sin x| ≤ |x|, with equality only when x = 0 .)
x→0

1.3: Finding Limits Analytically


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:
(a) lim f (x) = 0
x→1

(b) lim g(x) = 0


x→1

(c) lim f (x)/g(x) = 2


x→1

What can be said about the relative sizes of f (x) and g(x) as x approaches 1?

Problems
Using:
lim f (x) = 6 lim f (x) = 9 (1.E.1)
x→9 x→6

lim g(x) = 3 lim g(x) = 3 (1.E.2)


x→9 x→6

evaluate the limits given in Exercises 6-13, where possible. If it is not possible to know, state so.
6. lim(f (x) + g(x))
x→9

1.E.2 https://math.libretexts.org/@go/page/9966
7. lim(3f (x)/g(x))
x→9

f (x)−2g(x)
8. lim ( g(x)
)
x→9

f (x)
9. lim ( 3−g(x)
)
x→6

10. lim g(f (x))


x→9

11. lim f (g(x))


x→6

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


x→6

13. lim f (x)g(x) − f 2 2


(x) + g (x)
x→6

Using
lim f (x) = 2 lim f (x) = 1 (1.E.3)
x→1 x→10

lim g(x) = 0 lim g(x) = π (1.E.4)


x→1 x→10

evaluate the limits given in Exercises 14-17, where possible. If it is not possible to know, state so.
14. lim f (x ) g(x)

x→1

15. lim cos(g(x))


x→10

16. lim f (x)g(x)


x→1

17. lim g(5f (x))


x→1

In Exercises 18-32, evaluate the given limit.


18. lim x 2
− 3x + 7
x→3

19. lim (
x−3

x+5
)
x→π

20. lim cos x sin x


x→π/4

21. lim ln x
x→0

22. lim 4 x −8x

x→3

23. lim csc x


x→π/6

24. lim ln(1 + x)


x→0

25. lim
x +3x+5

5 x2 −2x−3
x→π

26. lim 3x+1

1−x
x→π

27. lim x −4x−12

x2 −13x+42
x→6

28. lim x +2x

x2 −2x
x→0

29. lim x +6x−16

x2 −3x+2
x→2

30. lim x −5x−14

x2 +10x+16
x→2

31. lim
x −5x−14
2
x +10x+16
x→−2

1.E.3 https://math.libretexts.org/@go/page/9966
2

32. lim
x +9x+8
2
x −6x−7
\
x→−1

Use the Squeeze Theorem in Exercises 33-36, where appropriate, to evaluate the given limit.
33. lim x sin( 1

x
)
x→0

34. lim sin x cos( 1


2
x
)
x→0

35. lim f (x), where 3x − 2 ≤ f (x) ≤ x 3


.
x→1

36. lim f (x), where 6x − 9 ≤ f (x) ≤ x on [0,3]. 2

x→3+

Exercises 37-40, challenge your understanding of limits but can be evaluated using the knowledge gained in this section.
37. lim sin 3x

x
x→0

38. lim sin 5x

8x
x→0

ln(1+x)
39. lim x
x→0

40. lim sin x

x
, where x is measured in degrees not radians.
x→0

1.4: One Sided Limits


Terms and Concepts
1. What are the three ways in which a limit may fail to exist?
2. T/F: If lim f (x) = 5 , then lim f (x) = 5
x→1− x→1

3. T/F: If lim f (x) = 5 , then lim f (x) = 5


x→1− x→1+

4. T/F: If lim f (x) = 5 , then lim f (x) = 5


x→1 x→1−

Problems
In Exercises 5-12, evaluate each expression using the given graph of f (x).
5.

1.E.4 https://math.libretexts.org/@go/page/9966
6.

7.

8.

9.

1.E.5 https://math.libretexts.org/@go/page/9966
10.

11.

12.

In Exercises 13-21, evaluate the given limits of the piecewise defined functions f .
x +1 x ≤1
13. f (x) = { 2
x −5 x >1

(a) lim f (x)



x→1

(b) lim f (x)


+
x→0

(c) lim f (x)


x→1

(d) f (1)
2
2x + 5x − 1 x <0
14. f (x) = {
sin x x ≥0

(a) lim f (x)



x→0

(b) lim f (x)


+
x→0

(c) lim f (x)


x→0

(d) f (0)

1.E.6 https://math.libretexts.org/@go/page/9966

⎧x 1 x < −1

15. f (x) = ⎨ x 3
+1 −1 ≤ x ≤ 1

2
x +1 x >1

(a) lim

f (x)
x→−1

(b) lim f (x)


+
x→1

(c) lim f (x)


x→−1

(d) f (−1)
(e) lim f (x)

x→1

(f) lim f (x)


+
x→1

(g) lim f (x)


x→1

(h) f (1)
cos x x <π
16. f (x) = {
sin x x ≥π

(a) lim f (x)



x→π

(b) lim f (x)


+
x→π

(c) lim f (x)


x→π

(d) f (π)
2
1 − cos x x <a
17. f (x) = { 2
, where a is a real number.
sin x x ≥a

(a) lim f (x)



x→a

(b) lim f (x)


+
x→a

(c) lim f (x)


x→a

(d) f (a)
⎧ x +1 x <1

18. f (x) = ⎨ 1 x =1

x −1 x >1

(a) lim f (x)



x→1

(b) lim f (x)


+
x→1

(c) lim f (x)


x→1

(d) f (1)
2
⎧x x <2

19. f (x) = ⎨ x + 1 x =2

2
−x + 2x + 4 x >2

(a) lim f (x)



x→2

(b) lim f (x)


+
x→2

(c) lim f (x)


x→2

(d) f (2)
2
a(x − b ) +c x <b
20. f (x) = { , where a, b and c are real numbers.
a(x − b) + c x ≥b

(a) lim f (x)



x→b

(b) lim f (x)


+
x→b

(c) lim f (x)


x→b

(d) f (b)

1.E.7 https://math.libretexts.org/@go/page/9966
|x|
x ≠0
21. f (x) = { x

0 x =0

(a) lim f (x)



x→0

(b) lim f (x)


+
x→0

(c) lim f (x)


x→0

(d) f (0)

Review
2

22. Evaluate the limit: lim


x +5x+4
2
x −3x−4
x→−1

23. Evaluate the limit: lim


x −16

x2 −4x−32
x→−4

24. Evaluate the limit: lim


x −15x+54
2
x −6x
x→−6

2
x −4.4x+1.6
25. Approximate the limit numerically: lim 2
x −0.4x
x→0.4

2
x +5.8x−1.2
26. Approximate the limit numerically: lim 2
x→0.2 x −4.2x+0.8

1.5: Continuity
Terms and Concepts
1. In your own words, describe what it means for a function to be continuous.
2. In your own words, describe what the Intermediate Value Theorem states.
3. What is a “root” of a function?
4. Given functions f and g on an interval I , how can the Bisection Method be used to find a value c where f (c) = g(c) ?
5. T/F: If f is defined on an open interval containing c, and lim f (x) exists, then f is continuous at c.
x→c

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


x→c

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


+
.
x→c

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



x→a

9. T/F: If f is continuous on [0, 1) and [1, 2), then f is continuous on [0, 2).
10. T/F: The sum of continuous functions is also continuous.

Problems
In Exercises 11-17, a graph of a function f is given along with a value a . Determine if f is continuous at a ; if it is not, state
why it is not.
11. a = 1

1.E.8 https://math.libretexts.org/@go/page/9966
12. a = 1

13. a = 1

14. a = 0

15. a = 1

16. a = 4

17.
(a) a = −2
(b) a = 0

1.E.9 https://math.libretexts.org/@go/page/9966
(c) a = 2

In Exercises 18-21, determine if f is continuous at the indicated values. If not, explain why.
1 x =0
18. f (x) = { sin x
x >0
x

(a) x = 0
(b) x = π
3
x −x x <1
19. f (x) = {
x −2 x ≥1

(a) x = 0
(b) x = 1
2
x +5x+4
x ≠ −1
20. f (x) = { x2 +3x+2

3 x = −1

(a) x = −1
(b) x = 10
2
x −64
x ≠8
21. f (x) = { 2
x −11x+24

5 x =8

(a) x = 0
(b) x = 8
In Exercises 22-32, give the intervals on which the given function is continuous.
22. f (x) = x 2
− 3x + 9

−−−−−
23. g(x) = √x2 − 4

−−−− −−−−
24. h(k) = √1 − k + √k + 1

−−− −−−−
25. f (t) = √5t 2
− 30

26. g(t) = 1

√1−t2

27. g(x) = 1+x


1
2

28. f (x) = e x

29. g(s) = ln s
30. h(t) = cos t
−−−−−
31. f (k) = √1 − e k

32. f (x) = sin(e x


+x )
2

33. Let f be continuous on [1,5] where f (1) = −2 and f (5) = −10 . Does a value 1 <c <5 exist such that f (c) = −9 ?
Why/why not?
34. Let g be continuous on [-3,7] where g(0) = 0 and g(2) = 25 . Does a value −3 < c < 7 exist such that g(c) = 15? Why/why
not?

1.E.10 https://math.libretexts.org/@go/page/9966
35. Let f be continuous on [-1,1] where f (−1) = −10 and f (1) = 10 . Does a value −1 < c < 1 exist such that f (c) = 11?

Why/why not?
36. Let h be continuous on [-1,1] where h(−1) = −10 and h(1) = 10 . Does a value −1 < c < 1 exist such that h(c) = 0?

Why/why not?
In Exercises 37-40, use the Bisection Method to approximate, accurate to two decimal places, the value of the root of the
given function in the given interval.
37. f (x) = x 2
+ 2x − 4 on [1, 1.5] .
38. f (x) = sin x − 1/2 on [0.5, 0.55].
39. f (x) = e x
− 2 on [0.65, 0.7] .
40. f (x) = cos x − sin x on [0.7, 0.8].

Review
2
x −5 x <5
41. Let f (x) = { .
5x x ≥5

(a) lim f (x)



x→5

(b) lim f (x)


+
x→5

(c) lim f (x)


x→5

(d) f (5)
42. Numerically approximate the following limits:
2

(a) lim
+
x −8.2x−7.2
2
x +5.8x+4
x→4/5
2

(b) lim

x −8.2x−7.2
2
x +5.8x+4
x→4/5

43. Give an example of function f (x) for which lim f (x) does not exist.
x→0

1.6: Limits Involving Infinity


Terms and Concepts
1. T/F: If lim f (x) = ∞ , then we are implicitly stating that the limit exists.
x→5

2. T/F: If lim f (x) = 5 , then we are implicitly stating that the limit exists.
x→∞

3. T/F: If lim f (x) = −∞



, then lim f (x) = ∞
+
.
x→1 x→1

4. T/F: If lim f (x) = ∞ , then f has a vertical asymptote at x = 5 .


x→5

5. T/F: ∞/0 is not an indeterminate form.


6. List 5 indeterminate forms.
7. Construct a function with a vertical asymptote at x = 5 and a horizontal asymptote at y = 5.
8. Let lim f (x) = ∞ . Explain how we know that f is/is not continuous at x = 7 .
x→7

Problems
In Exercises 9-14, evaluate the given limits using the graph of the function.
9. f (x) = 1
2
(x+1)

(a) lim

f (x)
x→−1

1.E.11 https://math.libretexts.org/@go/page/9966
(b) lim
+
f (x)
x→−1

10. f (x) = 1
2
(x−3)(x−5)

(a) lim f (x)



x→3

(b) lim f (x)


+
x→3

(c) lim f (x)


x→3

(d) lim f (x)



x→5

(e) lim f (x)


+
x→5

(f) lim f (x)


x→5

11. f (x) = 1

ex +1

(a) lim f (x)


x→−∞

(b) lim f (x)


x→∞

(c) lim f (x)



x→0

(d) lim f (x)


+
x→0

12. f (x) = x sin(πx) 2

(a) lim f (x)


x→−∞

(b) lim f (x)


x→∞

13. f (x) = cos(x)


(a) lim f (x)
x→−∞

1.E.12 https://math.libretexts.org/@go/page/9966
(b) lim f (x)
x→∞

14. f (x) = 2 + 10 x

(a) lim f (x)


x→−∞

(b) lim f (x)


x→∞

In Exercises 15-18, numerically approximate the following limits:


(a) lim f (x)

x→3

(b) lim f (x)


+
x→3

(c) lim f (x)


x→3

2
x −1
15. f (x) = 2
x −x−6

16. f (x) = 3
x +5x−36

x −5 x +3x+9
2

17. f (x) = 3
x −11x+30

x −4 x −3x+18
2

2
x −9x+18
18. f (x) = 2
x −x−6

In Exercises 19-24, identify the horizontal and vertical asymptotes, if any, of the given function.
2
2 x −2x−4
19. f (x) = 2
x +x−20

20. f (x) = −3 x −9x−6

5 x2 −10x−15

21. f (x) = 3
x +2−12

7 x −14 x −21x
2

22. f (x) = x −9

9x−9

2
x −9
23. f (x) = 9x+27

2
x −1
24. f (x) = −x −1
2

In Exercises 25-28, evaluate the given limit.


3 2
x +2 x +1
25. lim
x−5
x→∞

3 2

26. lim
x +2 x +1

5−x
x→∞

3 2

27. lim
x +2 x +1

x2 −5
x→∞

1.E.13 https://math.libretexts.org/@go/page/9966
3 2

28. lim
x +2 x +1

5−x
2
x→∞

Review
29. Use an ε − δ proof to show that lim 5x − 2 = 3 .
x→1

30. Let lim f (x) = 3 and lim g(x) = −1 . Evaluate the following limits.
x→2 x→2

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


x→2

(b) lim(f g)(x)


x→2

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


x→2

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

x→2

2
x −1 x <3
31. Let f (x) = { . Is f continuous everywhere?
x +5 x ≥3

32. Evaluate the limit: lim ln x .


x→c

1.E: Applications of Limits (Exercises) is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by LibreTexts.

1.E.14 https://math.libretexts.org/@go/page/9966
CHAPTER OVERVIEW
2: Derivatives
The previous chapter introduced the most fundamental of calculus topics: the limit. This chapter introduces the second most
fundamental of calculus topics: the derivative. Limits describe where a function is going; derivatives describe how fast the function
is going.

Topic hierarchy
2.1: Instantaneous Rates of Change- The Derivative
2.2: Interpretations of the Derivative
2.3: Basic Differentiation Rules
2.4: The Product and Quotient Rules
2.5: The Chain Rule
2.6: Implicit Differentiation
2.7: Derivatives of Inverse Functions
2.E: Applications of Derivatives(Exercises)

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 2: Derivatives is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et al. via
source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

1
2.1: Instantaneous Rates of Change- The Derivative
A common amusement park ride lifts riders to a height then allows them to freefall a certain distance before safely stopping them.
Suppose such a ride drops riders from a height of 150 feet. Students of physics may recall that the height (in feet) of the riders, t
seconds after freefall (and ignoring air resistance, etc.) can be accurately modeled by f (t) = −16t + 150 . 2


−−
Using this formula, it is easy to verify that, without intervention, the riders will hit the ground at t = 2.5√1.5 ≈ 3.06 seconds.
Suppose the designers of the ride decide to begin slowing the riders' fall after 2 seconds (corresponding to a height of 86 ft.). How
fast will the riders be traveling at that time?
We have been given a position function, but what we want to compute is a velocity at a specific point in time, i.e., we want an
instantaneous velocity. We do not currently know how to calculate this.
However, we do know from common experience how to calculate an average velocity. (If we travel 60 miles in 2 hours, we know
we had an average velocity of 30 mph.) We looked at this concept in Section 1.1 when we introduced the difference quotient. We
have
change in distance "rise''
= = average velocity. (2.1.1)
change in time run

We can approximate the instantaneous velocity at t = 2 by considering the average velocity over some time period containing
t = 2 . If we make the time interval small, we will get a good approximation. (This fact is commonly used. For instance, high speed

cameras are used to track fast moving objects. Distances are measured over a fixed number of frames to generate an accurate
approximation of the velocity.)
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) f (3) − f (2)
= = −80 ft/s, (2.1.2)
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 instantaneous velocity. On [2, 2.5] we have
f (2.5) − f (2) f (2.5) − f (2)
= = −72 ft/s. (2.1.3)
2.5 − 2 0.5

We can do this for smaller and smaller intervals of time. For instance, over a time span of 1/10 th
of a second, i.e., on , we
[2, 2.1]

have
f (2.1) − f (2) f (2.1) − f (2)
= = −65.6 ft/s. (2.1.4)
2.1 − 2 0.1

Over a time span of 1/100 th


of a second, on [2, 2.01], the average velocity is
f (2.01) − f (2) f (2.01) − f (2)
= = −64.16 ft/s. (2.1.5)
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)
(2.1.6)
h

where h is small.
What we really want is for h =0 , but this, of course, returns the familiar "0/0'' indeterminate form. So we employ a limit, as we
did in Section 1.1.
We can approximate the value of this limit numerically with small values of h as seen in Figure 2.1. It looks as though the velocity
is approaching −64 ft/s. Computing the limit directly gives

2.1.1 https://math.libretexts.org/@go/page/4157
2 2
f (2 + h) − f (2) −16(2 + h ) + 150 − (−16(2 ) + 150)
lim = lim
h→0 h h→0 h
2
−64h − 16h
= lim
h→0 h

= lim −64 − 16h


h→0

= −64.

Figure 2.1: Approximating the instantaneous velocity with average velocities over a small time period h .
Graphically, we can view the average velocities we computed numerically as the slopes of secant lines on the graph of f going
through the points (2, f (2)) and (2 + h, f (2 + h)) . In Figure 2.2, the secant line corresponding to h = 1 is shown in three
contexts. Figure 2.2(a) shows a "zoomed out'' version of f with its secant line. In (b), we zoom in around the points of intersection
between f and the secant line. Notice how well this secant line approximates f between those two points -- it is a common practice
to approximate functions with straight lines.
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 parts (c) and (d) of Figure 2.2, we zoom in around the point (2, 86). In (c) we see the secant line, which approximates f well, but
not as well the tangent line shown in (d).

Figure 2.2: Parts (a), (b) and (c) show the secant line to f (x) with h = 1, zoomed in different amounts. Part (d) shows the tangent
line to f at x = 2.
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.

2.1.2 https://math.libretexts.org/@go/page/4157
Definition 7: Derivative at a Point
Let f be a continuous function on an open interval I and let c be in I . The derivative of f at c , denoted f ′
(c) , is
f (c + h) − f (c)
lim , (2.1.7)
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 .

Definition 8: Tangent Line


Let fbe continuous on an open interval I and differentiable at c , for some c in I . The line with equation

is the tangent line to the graph of f at c ; that is, it is the line through (c, f (c)) whose slope is the
ℓ(x) = f (c)(x − c) + f (c)

derivative of f at c .

Some examples will help us understand these definitions.

Example 32: Finding derivatives and tangent lines

Let f (x) = 3x 2
+ 5x − 7 . Find:
1. f (1)

2. The equation of the tangent line to the graph of f at x = 1 .


3. f (3)

4. The equation of the tangent line to the graph f at x = 3 .


Solution
1. We compute this directly using Definition 7.
f (1 + h) − f (1)

f (1) = lim
h→0 h
2 2
3(1 + h ) + 5(1 + h) − 7 − (3(1 ) + 5(1) − 7)
= lim
h→0 h
2
3h + 11h
= lim
h→0 h

= lim 3h + 11 = 11.
h→0

2. 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 .


3. Again, using the definition,
f (3 + h) − f (3)

f (3) = lim
h→0 h
2 2
3(3 + h ) + 5(3 + h) − 7 − (3(3 ) + 5(3) − 7)
= lim
h→0 h
2
3h + 23h
= lim
h→0 h

= lim 3h + 23
h→0

= 23.

4. The tangent line at x = 3 has slope 23 and goes through the point (3, f (3)) = (3, 35). Thus the tangent line has equation
y = 23(x − 3) + 35 = 23x − 34 .

A graph of f (x) = 3x 2
+ 5x − 7 and its tangent lines at x = 1 and x = 3 .

2.1.3 https://math.libretexts.org/@go/page/4157
Figure 2.3: A graph of f (x) = 3x 2
+ 5x − 7 and its tangent lines at x = 1 and x = 3 .

Another important line that can be created using information from the derivative is the normal line. It is perpendicular to the
tangent line, hence its slope is the opposite--reciprocal of the tangent line's slope.

Definition 9: Normal Line


Let f be continuous on an open interval I and differentiable at c , for some c in I . The normal line to the graph of f at c is the
line with equation
−1
n(x) = (x − c) + f (c), (2.1.8)

f (c)

where f ′
(c) ≠ 0 . When f ′
(c) = 0 , the normal line is the vertical line through (c, f (c)); that is, x = c .

Example 33: Finding equations of normal lines


Let f (x) = 3x 2
+ 5x − 7 , as in Example 32. Find the equations of the normal lines to the graph of f at x = 1 and x = 3 .

Figure 2.4: A graph of f (x) = 3x 2


+ 5x − 7 , along with its normal line at x = 1 .
Solution
In Example 32, we found that f ′
(1) = 11 . Hence at x = 1 , the normal line will have slope −1/11. An equation for the normal
line is
−1
n(x) = (x − 1) + 1. (2.1.9)
11

The normal line is plotted with y = f (x) in Figure 2.4. Note how the line looks perpendicular to f . (A key word here is
"looks.'' Mathematically, we say that the normal line is perpendicular to f at x = 1 as the slope of the normal line is the
opposite--reciprocal of the slope of the tangent line. However, normal lines may not always look perpendicular. The aspect
ratio of the picture of the graph plays a big role in this.)
We also found that ′
f (3) = 23 , so the normal line to the graph of f at x =3 will have slope −1/23 . An equation for the
normal line is
−1
n(x) = (x − 3) + 35. (2.1.10)
23

2.1.4 https://math.libretexts.org/@go/page/4157
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.

Example 34: Finding the Derivative of a Line


Consider f (x) = 3x + 5 . Find the equation of the tangent line to f at x = 1 and x = 7 .
Solution
We find the slope of the tangent line by using Definition 7.
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 just found that f (1) = 3 . That is, we found the instantaneous rate of change of f (x) = 3x + 5 is 3. This is not surprising;

lines are characterized by being the only functions with a constant rate of change. That rate of change is called the slope of the
line. Since their rates of change are constant, their instantaneous rates of change are always the same; they are all the slope.
So given a line f (x) = ax + b , the derivative at any point x will be a ; that is, f ′
(x) = a .
It is now easy to see that the tangent line to the graph of f at x = 1 is just f , with the same being true for x = 7 .

We often desire to find the tangent line to the graph of a function without knowing the actual derivative of the function. In these
cases, the best we may be able to do is approximate the tangent line. We demonstrate this in the next example.

Example 35: Numerical Approximation of the Tangent Line

Approximate the equation of the tangent line to the graph of f (x) = sin x at x = 0 .

Figure 2.5: f (x) = sin x graphed with an approximation to its tangent line at x = 0 .
Solution
In order to find the equation of the tangent line, we need a slope and a point. The point is given to us: (0, sin 0) = (0, 0). To
compute the slope, we need the derivative. This is where we will make an approximation. Recall that
sin(0 + h) − sin 0

f (0) ≈ (2.1.11)
h

2.1.5 https://math.libretexts.org/@go/page/4157
for a small value of h . We choose (somewhat arbitrarily) to let h = 0.1 . Thus
sin(0.1) − sin 0

f (0) ≈ ≈ 0.9983. (2.1.12)
0.1

Thus our approximation of the equation of the tangent line is y = 0.9983(x − 0) + 0 = 0.9983x; it is graphed in Figure 2.5.
The graph seems to imply the approximation is rather good.

Recall from Section 1.3 that lim x→0 = 1 , meaning for values of x near 0, sin x ≈ x . Since the slope of the line y = x is 1 at
sin x

x = 0 , it should seem reasonable that "the slope of f (x) = sin x '' is near 1 at x = 0 . In fact, since we approximated the value of

the slope to be 0.9983, we might guess the actual value is 1. We'll come back to this later.
Consider again Example 32. To find the derivative of f at x = 1 , we needed to evaluate a limit. To find the derivative of f at
x = 3 , we needed to again evaluate a limit. We have this process:

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:

The output is the "derivative function,'' f ′


. The f
(x)

(x) function will take a number c as input and return the derivative of f at c .
This calls for a definition.

Definition 10: Derivative Function


Let f be a differentiable function on an open interval I . The function
f (x + h) − f (x)

f (x) = lim (2.1.13)
h→0 h

is the derivative of f .
Notation:
Let y = f (x). The following notations all represent the derivative:
dy df d d
′ ′
f (x) = y = = = (f ) = (y). (2.1.14)
dx dx dx dx

dy
Important: The notation is one symbol; it is not the fraction "dy/dx''. The notation, while somewhat confusing at first, was
dx

chosen with care. A fraction--looking symbol was chosen because the derivative has many fraction-like properties. 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.

Example 36: Finding the derivative of a function


Let f (x) = 3x 2
+ 5x − 7 as in Example 32. Find f ′
(x).}
Solution: We apply Definition 10.

2.1.6 https://math.libretexts.org/@go/page/4157
f (x + h) − f (x)

f (x) = lim
h→0 h
2 2
3(x + h ) + 5(x + h) − 7 − (3 x + 5x − 7)
= lim
h→0 h
2
3h + 6xh + 5h
= lim
h→0 h

= lim 3h + 6x + 5
h→0

= 6x + 5

So f (x) = 6x + 5 . Recall earlier we found that f


′ ′
(1) = 11 and f ′
(3) = 23 . Note our new computation of f ′
(x) affirm these
facts.

Example 37: Finding the derivative of a function


Let f (x) = 1

x+1
. Find f ′
.
(x)

Solution: We apply Definition 10.


f (x + h) − f (x)

f (x) = lim (2.1.15)
h→0 h
1 1

x+h+1 x+1
= lim
h→0 h

Now find common denominator then subtract; pull 1/h out front to facilitate reading.
1 x +1 x +h +1
= lim ⋅( − )
h→0 h (x + 1)(x + h + 1) (x + 1)(x + h + 1)

1 x + 1 − (x + h + 1)
= lim ⋅( )
h→0 h (x + 1)(x + h + 1)

1 −h
= lim ⋅( )
h→0 h (x + 1)(x + h + 1)

−1
= lim
h→0 (x + 1)(x + h + 1)

−1
=
(x + 1)(x + 1)

−1
=
2
(x + 1)

So f ′
(x) =
−1
2
. To practice using our notation, we could also state
(x+1)

d 1 −1
( ) = . (2.1.16)
2
dx x +1 (x + 1)

Example 38: Finding the derivative of a function


Find the derivative of f (x) = sin x .}
Solution
Before applying Definition 10, note that once this is found, we can find the actual tangent line to f (x) = sin x at x =0 ,
whereas we settled for an approximation in Example 35.

2.1.7 https://math.libretexts.org/@go/page/4157
We have found that when f (x) = sin x , f (x) = cos x. This should be somewhat surprising; the result of a tedious limit

process and the sine function is a nice function. Then again, perhaps this is not entirely surprising. The sine function is periodic
-- it repeats itself on regular intervals. Therefore its rate of change also repeats itself on the same regular intervals. We should
have known the derivative would be periodic; we now know exactly which periodic function it is.
Thinking back to Example 35, we can find the slope of the tangent line to f (x) = sin x at x =0 using our derivative. We
approximated the slope as 0.9983; we now know the slope is exactly cos 0 = 1 .

Example 39: Finding the derivative of a piecewise defined function


Find the derivative of the absolute value function,
−x x <0
f (x) = |x| = { . (2.1.17)
x x ≥0

See Figure 2.6.


Solution
f (x+h)−f (x)
We need to evaluate limh→0
h
. As f is piecewise--defined, we need to consider separately the limits when x <0

and when x > 0 .

Figure 2.6: The absolute value function, f (x) = |x| . Notice how the slope of the lines (and hence the tangent lines) abruptly
changes at x = 0 .
When x < 0 :
d −(x + h) − (−x)
( − x) = lim
dx h→0 h

−h
= lim
h→0 h

= lim −1
h→0

= −1.

2.1.8 https://math.libretexts.org/@go/page/4157
When x > 0 , a similar computation shows that d

dx
(x) = 1 .
We need to also find the derivative at x = 0 . By the definition of the derivative at a point, we have
f (0 + h) − f (0)

f (0) = lim . (2.1.18)
h→0 h

Since x = 0 is the point where our function's definition switches from one piece to 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.

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 differentiable at 0. So we have


−1 x <0
f (x) = { . (2.1.19)
1 x >0

At x = 0 , f ′
(x) does not exist; there is a jump discontinuity at 0; see Figure 2.7. So f (x) = |x| is differentiable everywhere
except at 0.

Figure 2.7: A graph of the derivative of f (x) = |x| .

The point of non-differentiability came where the piecewise defined function switched from one piece to the other. Our next
example shows that this does not always cause trouble.

Example 40: Finding the derivative of a piecewise defined function


sin x x ≤ π/2
Find the derivative of f (x), where f (x) = { . See Figure 2.8.
1 x > π/2

Figure 2.8: A graph of f (x) as defined in Example 40.


Solution

2.1.9 https://math.libretexts.org/@go/page/4157
Using Example 38, we know that when x < π/2, f ′
(x) = cos x . It is easy to verify that when x > π/2, f ′
(x) = 0 ; consider:
f (x + h) − f (x) 1 −1
lim = lim = lim 0 = 0. (2.1.20)
h→0 h h→0 h h→0

So far we have
cos x x < π/2

f (x) = { . (2.1.21)
0 x > π/2

We still need to find f ′


. Notice at x = π/2 that both pieces of f are 0, meaning we can state that f
(π/2)
′ ′
(π/2) = 0 .
Being more rigorous, we can again evaluate the difference quotient limit at x = π/2, utilizing again left and right--hand limits:

Since both the left and right hand limits are 0 at x = π/2, the limit exists and f ′
(π/2) exists (and is 0). Therefore we can fully
write f as


cos x x ≤ π/2
f (x) = { . (2.1.22)
0 x > π/2

See Figure 2.9 for a graph of this function.

Figure 2.9: A graph of f ′


(x) in Example 40.

Recall we pseudo--defined a continuous function as one in which we could sketch its graph without lifting our pencil. We can give
a pseudo--definition for differentiability as well: it is a continuous function that does not have any "sharp corners.'' One such sharp
corner is shown in Figure 2.6. Even though the function f in Example 40 is piecewise--defined, the transition is "smooth'' hence it
is differentiable. Note how in the graph of f in Figure 2.8 it is difficult to tell when f switches from one piece to the other; there is
no "corner.''
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?''

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

2.1.10 https://math.libretexts.org/@go/page/4157
This page titled 2.1: Instantaneous Rates of Change- The Derivative is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or
curated by Gregory Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history
is available upon request.

2.1.11 https://math.libretexts.org/@go/page/4157
2.2: Interpretations of the Derivative
The previous section defined the derivative of a function and gave examples of how to compute it using its definition (i.e., using
limits). The section also started with a brief motivation for this definition, that is, finding the instantaneous velocity of a falling
object given its position function. The next section will give us more accessible tools for computing the derivative, tools that are
easier to use than repeated use of limits.
This section falls in between the "What is the definition of the derivative?'' and "How do I compute the derivative?'' sections. Here
we are concerned with "What does the derivative mean?'', or perhaps, when read with the right emphasis, "What is the derivative?''
We offer two interconnected interpretations of the derivative, hopefully explaining why we care about it and why it is worthy of
study.

Interpretation of the Derivative #1: Instantaneous Rate of Change


The previous section started with an example of using the position of an object (in this case, a falling amusement--park rider) to
find the object's velocity. This type of example is often used when introducing the derivative because we tend to readily recognize
that velocity is the instantaneous rate of change of position. In general, if f is a function of x, then f (x) measures the

instantaneous rate of change of f with respect to x. Put another way, the derivative answers "When x changes, at what rate does f
change?'' Thinking back to the amusement--park ride, we asked "When time changed, at what rate did the height change?'' and
found the answer to be "By -64 feet per second.''
Now imagine driving a car and looking at the speedometer, which reads "60 mph.'' Five minutes later, you wonder how far you
have traveled. Certainly, lots of things could have happened in those 5 minutes; you could have intentionally sped up significantly,
you might have come to a complete stop, you might have slowed to 20 mph as you passed through construction. But suppose that
you know, as the driver, none of these things happened. You know you maintained a fairly consistent speed over those 5 minutes.
What is a good approximation of the distance traveled?
One could argue the only good approximation, given the information provided, would be based on "distance = rate × time.'' In this
case, we assume a constant rate of 60 mph with a time of 5/60 hours. Hence we would approximate the distance traveled as 5
miles.
Referring back to the falling amusement--park ride, knowing that at t = 2 the velocity was −64 ft/s, we could reasonably assume
that 1 second later the riders' height would have dropped by about 64 feet. Knowing that the riders were accelerating as they fell
would inform us that this is an under--approximation. If all we knew was that f (2) = 86 and f (2) = −64 , we'd know that we'd

have to stop the riders quickly otherwise they would hit the ground!

Units of the Derivative


It is useful to recognize the units of the derivative function. If y is a function of x, i.e., y = f (x) for some function f , and y is
measured in feet and x in seconds, then the units of y = f are "feet per second,'' commonly written as "ft/s.'' In general, if y is
′ ′

measured in units P and x is measured in units Q, then y will be measured in units "P per Q'', or "P /Q.'' Here we see the

fraction--like behavior of the derivative in the notation:


dy units of y
the units of are . (2.2.1)
dx units of x

Example 41: The meaning of the derivative: World Population


Let represent the world population t minutes after 12:00 a.m., January 1, 2012. It is fairly accurate to say that
P (t)

P (0) = 7, 028, 734, 178 (www.prb.org). 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, 800people.

Example 42: The meaning of the derivative: Manufacturing


The term widget is an economic term for a generic unit of manufacturing output. Suppose a company produces widgets and
knows that the market supports a price of $10 per widget. Let P (n) give the profit, in dollars, earned by manufacturing and

2.2.1 https://math.libretexts.org/@go/page/4158
selling n widgets. The company likely cannot make a (positive) profit making just one widget; the start--up costs will likely
exceed $10. Mathematically, we would write this as P (1) < 0 .
What do P (1000) = 500 and P ′
(1000) = 0.25 mean? Approximate P (1100).
Solution:
The equation P (1000) = 500 means that selling 1,000 widgets returns a profit of \$500. We interpret P (1000) = 0.25 as ′

meaning that the profit is increasing at rate of $0.25 per widget (the units are "dollars per widget.'') Since we have no other
information to use, our best approximation for P (1100) is:

P (1100) ≈ P (1000) + P (1000) × 100 = $500 + 100 ⋅ 0.25 = $525. (2.2.2)

We approximate that selling 1,100 widgets returns a profit of $525.

The previous examples made use of an important approximation tool that we first used in our previous "driving a car at 60 mph''
example at the beginning of this section. Five minutes after looking at the speedometer, our best approximation for distance
traveled assumed the rate of change was constant. In Examples 41 and 42 we made similar approximations. We were given rate of
change information which we used to approximate total change. Notationally, we would say that

f (c + h) ≈ f (c) + f (c) ⋅ h. (2.2.3)

This approximation is best when h is "small.'' "Small'' is a relative term; when dealing with the world population, h = 22 days =
28,800 minutes is small in comparison to years. When manufacturing widgets, 100 widgets is small when one plans to manufacture
thousands.

The Derivative and Motion


One of the most fundamental applications of the derivative is the study of motion. Let s(t) be a position function, where t is time
and s(t) is distance. For instance, s could measure the height of a projectile or the distance an object has traveled.
Let's let s(t) measure the distance traveled, in feet, of an object after t seconds of travel. Then ′
s (t) has units "feet per second,''
and s (t) measures the instantaneous rate of distance change -- it measures velocity.

Now consider v(t) , a velocity function. That is, at time t , v(t) gives the velocity of an object. The derivative of v , v (t) , gives the

instantaneous rate of velocity change -- acceleration. (We often think of acceleration in terms of cars: a car may "go from 0 to 60
in 4.8 seconds.'' This is an average acceleration, a measurement of how quickly the velocity changed.) If velocity is measured in
feet per second, and time is measured in seconds, then the units of acceleration (i.e., the units of v (t) ) are "feet per second per

second,'' or ((ft/s)/s). We often shorten this to "feet per second squared,'' or ft/s , but this tends to obscure the meaning of the
2

units.
Perhaps the most well known acceleration is that of gravity. In this text, we use 2
g = 32ft/s or g = 9.8m/s
2
. What do these
numbers mean?
A constant acceleration of 32(ft/s)/s means that the velocity changes by 32ft/s each second. For instance, let v(t) measures the
velocity of a ball thrown straight up into the air, where v has units ft/s and t is measured in seconds. The ball will have a positive
2
velocity while traveling upwards and a negative velocity while falling down. The acceleration is thus −32ft/s . If v(1) = 20 ft/s,
then when t = 2 , the velocity will have decreased by 32ft/s; that is, v(2) = −12 ft/s. We can continue: v(3) = −44 ft/s, and we can
also figure that v(0) = 52 ft/s.
These ideas are so important we write them out as a Key Idea.

Key Idea 1: The Derivative and Motion


1. Let s(t) be the position function of an object. Then s (t) is the velocity function of the object.

2. Let v(t) be the velocity function of an object. Then v (t) is the acceleration function of the object.

We now consider the second interpretation of the derivative given in this section. This interpretation is not independent from the
first by any means; many of the same concepts will be stressed, just from a slightly different perspective.

2.2.2 https://math.libretexts.org/@go/page/4158
Interpretation of the Derivative #2: The Slope of the Tangent Line
f (c+h)−f (c)
Given a function y = f (x), the difference quotient h
gives a change in y values divided by a change in x values; i.e., it is
a measure of the "rise over run,'' or "slope,'' of the line that goes through two points on the graph of f : (c, f (c)) and
(c + h, f (c + h)) . As h shrinks to 0, these two points come close together; in the limit we find f (c), the slope of a special line

called the tangent line that intersects f only once near x = c .


Lines have a constant rate of change, their slope. Nonlinear functions do not have a constant rate of change, but we can measure
their instantaneous rate of change at a given x value c by computing f (c). We can get an idea of how f is behaving by looking at

the slopes of its tangent lines. We explore this idea in the following example.

Example 43: Understanding the derivative: the rate of change


Consider f (x) = x as shown in Figure 2.10. It is clear that at
2
x =3 the function is growing faster than at x =1 , as it is
steeper at x = 3 . How much faster is it growing?
Solution:
We can answer this directly after the following section, where we learn to quickly compute derivatives. For now, we will
answer graphically, by considering the slopes of the respective tangent lines.

Figure 2.10: A graph of f (x) = x . 2

With practice, one can fairly effectively sketch tangent lines to a curve at a particular point. In Figure 2.11, we have sketched
the tangent lines to f at x = 1 and x = 3 , along with a grid to help us measure the slopes of these lines. At x = 1 , the slope is
2; at x = 3 , the slope is 6. Thus we can say not only is f growing faster at x = 3 than at x = 1 , it is growing three times as
fast.

Figure 2.11: A graph of f (x) = x and tangent lines.


2

Example 44: Understanding the graph of the derivative


Consider the graph of f (x) and its derivative, f ′
(x), in Figure 2.12(a). Use these graphs to find the slopes of the tangent lines
to the graph of f at x = 1 , x = 2 , and x = 3 .

2.2.3 https://math.libretexts.org/@go/page/4158
Figure 2.12: Graphs of f and f in Example 44, along with tangent lines in (b).

Solution
To find the appropriate slopes of tangent lines to the graph of f , we need to look at the corresponding values of f . ′

The slope of the tangent line to f at x = 1 is f (1); this looks to be about −1.

The slope of the tangent line to f at x = 2 is f (2); this looks to be about 4.


The slope of the tangent line to f at x = 3 is f (3); this looks to be about 3.


Using these slopes, the tangent lines to f are sketched in Figure 2.12(b). Included on the graph of f

in this figure are filled
circles where x = 1 , x = 2 and x = 3 to help better visualize the y value of f at those points.

Example 45: Approximation with the derivative


Consider again the graph of f (x) and its derivative f ′
(x) in Example 44. Use the tangent line to f at x = 3 to approximate the
value of f (3.1).
Solution
Figure 2.13 shows the graph of f along with its tangent line, zoomed in at x = 3 . Notice that near x = 3 , the tangent line
makes an excellent approximation of f . Since lines are easy to deal with, often 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.)

Figure 2.13: Zooming in on f at x = 3 for the function given in Examples 44 and 45.
While the tangent line to f was drawn in Example 44, it was not explicitly computed. Recall that the tangent line to f at x = c
is y = f (c)(x − c) + f (c) . While f is not explicitly given, by the graph it looks like f (3) = 4 . Recalling that f (3) = 3 , we
′ ′

can compute the tangent line to be approximately y = 3(x − 3) + 4. It is often useful to leave the tangent line in point--slope
form.
To use the tangent line to approximate f (3.1), we simply evaluate y at 3.1 instead of f .
f (3.1) ≈ y(3.1) = 3(3.1 − 3) + 4 = .1 ∗ 3 + 4 = 4.3. (2.2.4)

2.2.4 https://math.libretexts.org/@go/page/4158
We approximate f (3.1) ≈ 4.3.

To demonstrate the accuracy of the tangent line approximation, we now state that in Example 45, f (x) = −x + 7x − 12x + 4 . 3 2

We can evaluate f (3.1) = 4.279. Had we known f all along, certainly we could have just made this computation. In reality, we
often only know two things:
1. What f (c) is, for some value of c , and
2. what f (c) is.

For instance, we can easily observe the location of an object and its instantaneous 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 explained above after Example 42. We stated there that

f (c + h) ≈ f (c) + f (c) ⋅ h. (2.2.5)

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 45, 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:
′ ′
y(c + h) = f (c)((c + h) − c) + f (c) = f (c) ⋅ h + f (c). (2.2.6)

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.

This page titled 2.2: Interpretations of the Derivative is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by
Gregory Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available
upon request.

2.2.5 https://math.libretexts.org/@go/page/4158
2.3: Basic Differentiation Rules
The derivative is a powerful tool but is admittedly awkward given its reliance on limits. Fortunately, one thing mathematicians are
good at is abstraction. For instance, instead of continually finding derivatives at a point, we abstracted and found the derivative
function.
Let's practice abstraction on linear functions, y = mx + b . What is y ? Without limits, recognize that linear function are

characterized by being functions with a constant rate of change (the slope). The derivative, y , gives the instantaneous rate of

change; with a linear function, this is constant, m. Thus y = m . ′

Let's abstract once more. Let's find the derivative of the general quadratic function, f (x) = ax 2
+ bx + c . Using the definition of
the derivative, we have:
2 2
a(x + h ) + b(x + h) + c − (ax + bx + c)

f (x) = lim
h→0 h
2
ah + 2ahx + bh
= lim
h→0 h

= lim ah + 2ax + b
h→0

= 2ax + b.

So if y = 6x 2
+ 11x − 13 , we can immediately compute y ′
= 12x + 11 .
In this section (and in some sections to follow) we will learn some of what mathematicians have already discovered about the
derivatives of certain functions and how derivatives interact with arithmetic operations. We start with a theorem.

Theorem 12: Derivatives of Common Functions

This theorem starts by stating an intuitive fact: constant functions have no 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 = x ) are very straightforward: multiply by the power, then subtract 1 from the
n

power. We see something incredible about the function y = e : it is its own derivative. We also see a new connection between the
x

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 0
f (x) = (x) = (x ) = 1 ⋅ x = 1. (2.3.1)
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?'' The answer is: 1. They change at the same rate.
Let's practice using this theorem.

Example 46: Using Theorem 12 to find, and use, derivatives


Let f (x) = x . 3

1. Find f (x).

2. Find the equation of the line tangent to the graph of f at x = −1 .


3. Use the tangent line to approximate (−1.1) . 3

2.3.1 https://math.libretexts.org/@go/page/4159
4. Sketch f , f and the found tangent line on the same axis.

Solution
1. The Power Rule states that if f (x) = x , then f (x) = 3x .
3 ′ 2

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 equation

y = 3(x − (−1)) + (−1) = 3x + 2 .

3. We can use the tangent line to approximate (−1.1) as −1.1 is close to −1. We have
3

3
(−1.1 ) ≈ 3(−1.1) + 2 = −1.3. (2.3.2)

We can easily find the actual answer; (−1.1) 3


= −1.331 .
4. See Figure 2.14.

Figure 2.14: A graph of f (x) = x , along with its derivatives f


3 ′
(x) = 3x
2
and its tangent line at x = −1 .

Theorem 12 gives useful information, but we will need much more. For instance, using the theorem, we can easily find the
derivative of y = x , but it does not tell how to compute the derivative of y = 2x , y = x + sin x nor y = x sin x . The
3 3 3 3

following theorem helps with the first two of these examples (the third is answered in the next section).

Theorem 13: Properties of the Derivative


Let f and g be differentiable on an open interval I and let c be a real number. Then:
1. Sum/Difference Rule:
d d d ′ ′
(f (x) ± g(x)) = (f (x)) ± (g(x)) = f (x) ± g (x)
dx dx dx

2. Constant Multiple Rule:


d

dx
(c ⋅ f (x)) = c ⋅
d

dx
(f (x)) = c ⋅ f (x)

.

Theorem 13 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 36 that we found, using the limit definition, the derivative of
f (x) = 3 x + 5x − 7 . We can now find its derivative without expressly using limits:
2

d 2
d 2
d d
(3 x + 5x + 7) =3 (x ) + 5 (x) + (7)
dx dx dx dx

= 3 ⋅ 2x + 5 ⋅ 1 + 0

= 6x + 5.

We were a bit pedantic here, showing every step. Normally we would do all the arithmetic and steps in our head and readily find
d 2
(3 x + 5x + 7) = 6x + 5.
dx

Example 47: Using the tangent line to approximate a function value

Let f (x) = sin x + 2x + 1 . Approximate f (3) using an appropriate tangent line.


Solution
This problem is intentionally ambiguous; we are to approximate using an appropriate tangent line. How good of an
approximation are we seeking? What does appropriate mean?

2.3.2 https://math.libretexts.org/@go/page/4159
In the "real world,'' people solving problems deal with these issues all time. One must make a judgment using whatever seems
reasonable. In this example, the actual answer is f (3) = sin 3 + 7 , where the real problem spot is sin 3 . What is sin 3 ?
Since 3 is close to π, we can assume sin 3 ≈ sin π = 0 . Thus one guess is f (3) ≈ 7 . Can we do better? Let's use a tangent line
as instructed and examine the results; it seems best to find the tangent line at x = π .
Using Theorem 12 we find f (x) = cos x + 2 . The slope of the tangent line is thus f (π) = cos π + 2 = 1 . Also,
′ ′

f (π) = 2π + 1 ≈ 7.28 . So the tangent line to the graph of f at x = π is y = 1(x − π) + 2π + 1 = x + π + 1 ≈ x + 4.14 .

Evaluated at x = 3 , our tangent line gives y = 3 + 4.14 = 7.14 . Using the tangent line, our final approximation is that
f (3) ≈ 7.14.

Using a calculator, we get an answer accurate to 4 places after the decimal: f (3) = 7.1411 . Our initial guess was 7; our
tangent line approximation was more accurate, at 7.14.
The point is not "Here's a cool way to do some math without a calculator.'' Sure, that might be handy sometime, but your phone
could probably give you the answer. Rather, the point is to say that tangent lines are a good way of approximating, and many
scientists, engineers and mathematicians often face problems too hard to solve directly. So they approximate.

Higher Order Derivatives


The derivative of a function f is itself a function, therefore we can take its derivative. The following definition gives a name to this
concept and introduces its notation.

Definition 11: Higher Order Derivatives

Let y = f (x) be a differentiable function on I .


1. The second derivative of f is:
2
′′
d ′
d dy d y ′′
f (x) = (f (x)) = ( ) = =y . (2.3.3)
2
dx dx dx dx

1. The third derivative of f is:


2 3
′′′
d ′′
d d y d y ′′′
f (x) = (f (x)) = ( ) = =y . (2.3.4)
2 3
dx dx dx dx

1. The n th
derivative} of f is:
n−1 n
(n)
d (n−1)
d d y d y (n)
f (x) = (f (x)) = ( ) = =y . (2.3.5)
n−1 n
dx dx dx dx

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
too confusing.
Let's practice using this new concept.

Example 48: Finding higher order derivatives

Find the first four derivatives of the following functions:


1. f (x) = 4x 2

2. f (x) = sin x
3. f (x) = 5e x

Solution
1. Using the Power and Constant Multiple Rules, we have: f ′
(x) = 8x . Continuing on, we have

′′
d ′′′ (4)
f (x) = (8x) = 8; f (x) = 0; f (x) = 0. (2.3.6)
dx

Notice how all successive derivatives will also be 0.


2. We employ Theorem 12 repeatedly.

2.3.3 https://math.libretexts.org/@go/page/4159
′ ′′ ′′′ (4)
f (x) = cos x; f (x) = − sin x; f (x) = − cos x; f (x) = sin x. (2.3.7)

Note how we have come right back to f (x) again. (Can you quickly figure what f (23)
(x) is?)
3. Employing Theorem 12 and the Constant Multiple Rule, we can see that
′ ′′ ′′′ (4) x
f (x) = f (x) = f (x) = f (x) = 5 e . (2.3.8)

Interpreting Higher Order Derivatives


What do higher order derivatives mean? What is the practical interpretation? Our first answer is a bit wordy, but is technically
correct and beneficial to understand. That is,
The second derivative of a function f is the rate of change of the rate of change of f . (2.3.9)

One way to grasp this concept is to let f describe a position function. Then, as stated in Key Idea 1, f describes the rate of ′

position change: velocity. We now consider f , which describes the rate of velocity change. Sports car enthusiasts talk of how fast
′′

a car can go from 0 to 60 mph; they are bragging about the acceleration of the car.
We started this chapter with amusement--park riders free--falling with position function f (t) = −16t + 150 . It is easy to compute 2

f (t) = −32t ft/s and f (t) = −32 (ft/s)/s. We may recognize this latter constant; it is the acceleration 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
shortened to "feet per second squared,'' written as "ft/s .'' 2

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.

This page titled 2.3: Basic Differentiation Rules is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory
Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

2.3.4 https://math.libretexts.org/@go/page/4159
2.4: The Product and Quotient Rules
The previous section showed that, in some ways, derivatives behave nicely. The Constant Multiple and Sum/Difference Rules
established that the derivative of f (x) = 5x + sin x was not complicated. We neglected computing the derivative of things like
2

g(x) = 5 x sin x and h(x) = on purpose; their derivatives are not as straightforward. (If you had to guess what their
2 5x

sin x

respective derivatives are, you would probably guess wrong.) For these, we need the Product and Quotient Rules, respectively,
which are defined in this section.
We begin with the Product Rule.

Theorem 14: Product Rule


Let f and g be differentiable functions on an open interval I . Then f g is a differentiable function on I , and
d ′ ′
(f (x)g(x)) = f (x)g (x) + f (x)g(x). (2.4.1)
dx

Important: d

dx
′ ′
(f (x)g(x)) ≠ f (x)g (x) ! While this answer is simpler than the Product Rule, it is wrong.

We practice using this new rule in an example, followed by an example that demonstrates why this theorem is true.

Example 49: Using the Product Rule


Use the Product Rule to compute the derivative of y = 5x 2
sin x . Evaluate the derivative at x = π/2.
Solution
To make our use of the Product Rule explicit, let's set f (x) = 5x and 2
g(x) = sin x . We easily compute/recall that
f (x) = 10x and g (x) = cos x . Employing the rule, we have
′ ′

d
2 2
(5 x sin x) = 5 x cos x + 10x sin x. (2.4.2)
dx

At x = π/2, we have
2

π π π π
y (π/2) = 5 ( ) cos( ) + 10 sin( ) = 5π. (2.4.3)
2 2 2 2

We graph y and its tangent line at x = π/2, which has a slope of 5π, in Figure 2.15. While this does not prove that the Product
Rule is the correct way to handle derivatives of products, it helps validate its truth.

Figure 2.15: A graph of y = 5x 2


sin x and its tangent line at x = π/2 .

We now investigate why the Product Rule is true.

Example 50: A proof of the Product Rule


Use the definition of the derivative to prove Theorem 14.
Solution
By the limit definition, we have

2.4.1 https://math.libretexts.org/@go/page/4160
d f (x + h)g(x + h) − f (x)g(x)
(f (x)g(x)) = lim . (2.4.4)
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 + h)g(x) + f (x + h)g(x) , then do some regrouping as shown.

d f (x + h)g(x + h) − f (x)g(x)
(f (x)g(x)) = lim (now add 0 to the numerator)
dx h→0 h

f (x + h)g(x + h) − f (x + h)g(x) + f (x + h)g(x) − f (x)g(x)


= lim (regroup)
h→0 h

(f (x + h)g(x + h) − f (x + h)g(x)) + (f (x + h)g(x) − f (x)g(x))

= lim
h→0 h

f (x + h)g(x + h) − f (x + h)g(x) f (x + h)g(x) − f (x)g(x)


= lim + lim (factor)
h→0 h h→0 h

g(x + h) − g(x) f (x + h) − f (x)


= lim f (x + h) + lim g(x) (apply limits)
h→0 h h→0 h
′ ′
= f (x)g (x) + f (x)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.''

Example 51: Exploring alternate derivative methods

Let y = (x + 3x + 1)(2x − 3x + 1) . Find y rime two ways: first, by expanding the given product and then taking the
2 2 p

derivative, and second, by applying the Product Rule. Verify that both methods give the same answer.
Solution
We first expand the expression for y ; a little algebra shows that y = 2x 4
+ 3x
3
− 6x
2
+1 . It is easy to compute y ;′

′ 3 2
y = 8x + 9x − 12x. (2.4.5)

Now apply the Product Rule.


′ 2 2
y = (x + 3x + 1)(4x − 3) + (2x + 3)(2 x − 3x + 1)

3 2 3
= (4 x + 9x − 5x − 3) + (4 x − 7x + 3)

3 2
= 8x + 9x − 12x.

The uninformed usually assume that "the derivative of the product is the product of the derivatives.'' Thus we are tempted to
say that y = (2x + 3)(4x − 3) = 8x + 6x − 9 . Obviously this is not correct.
′ 2

Example 52: Using the Product Rule with a product of these three functions
Let y = x 3
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 together, and consider y = x ( ln x cos x). 3

Following the Product Rule, we have


′ 3 ′ 2
y = (x )( ln x cos x ) + 3 x ( ln x cos x)


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

3
1 2
= (x )( ln x(− sin x) + cos x) + 3 x ( ln x cos x)
x

3 3
1 2
=x ln x(− sin x) + x cos x + 3 x ln x cos x
x

2.4.2 https://math.libretexts.org/@go/page/4160
Recognize the pattern in our answer above: when applying the Product Rule to a product of three functions, there are three
terms added together in the final derivative. Each terms 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 4 or more functions.

We consider one more example before discussing another derivative rule.

Example 53: Using the Product Rule


Find the derivatives of the following functions.
1. f (x) = x ln x
2. g(x) = x ln x − x .
Solution:
Recalling that the derivative of ln x is 1/x, we use the Product Rule to find our answers.

1. d

dx
(x ln x) = x ⋅ 1/x + 1 ⋅ ln x = 1 + ln x .
2. Using the result from above, we compute
d
(x ln x − x) = 1 + ln x − 1 = ln x. (2.4.6)
dx

This seems significant; if the natural log function ln x is an important function (it is), it seems worthwhile to know a function
whose derivative is ln x. We have found one. (We leave it to the reader to find another; a correct answer will be very similar to
this one.)

We have learned how to compute the derivatives of sums, differences, and products of functions. We now learn how to find the
derivative of a quotient of functions.

Theorem 15: Quotient Rule


Let f and g be functions defined on an open interval I , where g(x) ≠ 0 on I . Then f /g is differentiable on I , and
′ ′
d f (x) g(x)f (x) − f (x)g (x)
( ) = . (2.4.7)
2
dx g(x) g(x)

The Quotient Rule is not hard to use, although it might be a bit tricky to remember. A useful mnemonic works as follows. Consider
a fraction's numerator and denominator as "HI'' and "LO'', respectively. Then
d HI LO ⋅ dHI − HI ⋅ dLO
( ) = , (2.4.8)
dx LO LOLO

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.

Example 54: Using the Quotient Rule


2

Let f (x) = 5x

sin x
. Find f ′
.
(x)

Solution
Directly applying the Quotient Rule gives:
2 2
d 5x sin x ⋅ 10x − 5 x ⋅ cos x
( ) =
2
dx sin x sin x
2
10x sin x − 5 x cos x
= .
2
sin x

2.4.3 https://math.libretexts.org/@go/page/4160
The Quotient Rule allows us to fill in holes in our understanding of derivatives of the common trigonometric functions. We start
with finding the derivative of the tangent function.

Example 55

Using the Quotient Rule to find dx


d
( tan x) .
Find the derivative of y = tan x .
Solution
At first, one might feel unequipped to answer this question. But recall that tan x = sin x/ cos x, so we can apply the Quotient
Rule.
d d sin x
( tan x) = ( )
dx dx cos x

cos x cos x − sin x(− sin x)


=
2
cos x
2 2
cos x + sin x
=
cos2 x
1
=
2
cos x
2
= sec x.

This is a beautiful result. To confirm its truth, we can find the equation of the tangent line to y = tan x at x = π/4. The slope
is sec (π/4) = 2 ; y = tan x , along with its tangent line, is graphed in Figure 2.16.
2

Figure 2.16: A graph of y = tan x along with its tangent line at x = π/4 .

We include this result in the following theorem about the derivatives of the trigonometric functions. Recall we found the derivative
of y = sin x in Example 38 and stated the derivative of the cosine function in Theorem 12. The derivatives of the cotangent,
cosecant and secant functions can all be computed directly using Theorem 12 and the Quotient Rule.

Theorem 16: Derivatives of Trigonometric Functions

To remember the above, it may be helpful to keep in mind that the derivatives of the trigonometric functions that start with "c'' have
a minus sign in them.

2.4.4 https://math.libretexts.org/@go/page/4160
Example 56: Exploring alternative derivative methods
2

In Example 54 the derivative of f (x) = was found using the Quotient Rule. Rewriting 5x

sin x
f as 2
f (x) = 5 x csc x , find f

using Theorem 16 and verify the two answers are the same.}
Solution
2

We found in Example 54 that the ′


f (x) =
10x sin x−5 x

2
cos x
. We now find f

using the Product Rule, considering f as
sin x

f (x) = 5 x
2
csc x .
d
′ 2
f (x) = (5 x csc x)
dx
2
= 5 x (− csc x cot x) + 10x csc x (now rewrite trig functions)

−1 cos x 10x
2
= 5x ⋅ ⋅ +
sin x sin x sin x
2
−5 x cos x 10x
= + (get common denominator)
2
sin x sin x
2
10x sin x − 5 x cos x
=
2
sin 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 show in the next example.

Example 57: Using the Quotient Rule to expand the Power Rule
Find the derivatives of the following functions.
1. f (x) = 1

2. f (x) = 1

x
n
, where n > 0 is an integer.
Solution
We employ the Quotient Rule.
1. f ′
(x) =
x⋅0−1⋅1
2
x
=−
1
2
x
.
n n−1 n−1
x ⋅0−1⋅n x
2. f ′
(x) = 2
=−
nx

x
2n
=−
x
n
n+1
.
(xn )

The derivative of y = 1
n
x
turned out to be rather nice. It gets better. Consider:
d 1 d −n
( ) = (x ) (apply result from Example 57)
n
dx x dx
n
=− (rewrite algebraically)
n+1
x
−(n+1)
= −nx
−n−1
= −nx .

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 restriction of n > 0 .

Theorem 17: Power Rule with Integer Exponents


Let f (x) = x , where n ≠ 0 is an integer. Then
n

′ n−1
f (x) = n ⋅ x . (2.4.9)

2.4.5 https://math.libretexts.org/@go/page/4160
Taking the derivative of many functions is relatively straightforward. It is clear (with practice) what rules apply and in what order
they should be applied. Other functions present multiple paths; different rules may be applied depending on how the function is
treated. One of the beautiful things about calculus is that there is not "the'' right way; each path, when applied correctly, leads to the
same result, the derivative. We demonstrate this concept in an example.

Example 58: Exploring alternate derivative methods


2

Let f (x) = x −3x+1

x
. Find f ′
(x) in each of the following ways:
1. By applying the Quotient Rule,
2. by viewing f as f (x) = (x − 3x + 1) ⋅ x2
and applying the Product and Power Rules, and
−1

3. by "simplifying\primeskip'' first through division.


Verify that all three methods give the same result.
Solution
1. Applying the Quotient Rule gives:
2 2
x ⋅ (2x − 3) − (x − 3x + 1) ⋅ 1 x −1 1

f (x) = = =1− . (2.4.10)
2 2 2
x x x

2. By rewriting f , we can apply the Product and Power Rules as follows:


′ 2 −2 −1
f (x) = (x − 3x + 1) ⋅ (−1)x + (2x − 3) ⋅ x

2
x − 3x + 1 2x − 3
=− +
2
x x
2 2
x − 3x + 1 2x − 3x
=− +
2 2
x x
2
x −1 1
= =1− ,
2 2
x x

the same result as above.


3. As x ≠ 0 , we can divide through by x first, giving f (x) = x − 3 + 1

x
. Now apply the Power Rule.
1

f (x) = 1 − , (2.4.11)
2
x

the same result as before.

Example 58 demonstrates three methods of finding f . One is hard pressed to argue for a "best method'' as all three gave the same

result without too much difficulty, although it is clear that using the Product Rule required more steps. Ultimately, the important
principle to take away from this is: reduce the answer to a form that seems "simple'' and easy to interpret. In that example, we saw
different expressions for f , including:

2
1 x ⋅ (2x − 3) − (x − 3x + 1) ⋅ 1
2 −2 −1
1− = = (x − 3x + 1) ⋅ (−1)x + (2x − 3) ⋅ x . (2.4.12)
2 2
x x

They are equal; they are all correct; only the first is "clear.'' Work to make answers clear.
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 memorize the derivatives of a certain set of functions, such as "the derivative of sin x is cos x.'' The
Sum/Difference, Constant Multiple, Power, Product and Quotient Rules show us how to find the derivatives of certain
combinations of these functions. The next section shows how to find the derivatives when we compose these functions together.

This page titled 2.4: The Product and Quotient Rules is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by
Gregory Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available
upon request.

2.4.6 https://math.libretexts.org/@go/page/4160
2.5: The Chain Rule
We have covered almost all of the derivative rules that deal with combinations of two (or more) functions. The operations of addition,
subtraction, multiplication (including by a constant) and division led to the Sum and Difference rules, the Constant Multiple Rule, the
Power Rule, the Product Rule and the Quotient Rule. To complete the list of differentiation rules, we look at the last way two (or
more) functions can be combined: the process of composition (i.e. one function "inside'' another).
One example of a composition of functions is f (x) = cos(x ) . We currently do not know how to compute this derivative. If forced to
2

guess, one would likely guess f (x) = − sin(2x),where we recognize − sin x as the derivative of cos x and 2x as the derivative of

x . However, this is not the case; f (x) ≠ − sin(2x) . In Example 62 we'll see the correct answer, which employs the new rule this
2 ′

section introduces, the Chain Rule.


Before we define this new rule, recall the notation for composition of functions. We write (f ∘ g)(x) or f (g(x)),read as "f of g of x,''
to denote composing f with g . In shorthand, we simply write f ∘ g or f (g) and read it as "f of g .'' Before giving the corresponding
differentiation rule, we note that the rule extends 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.

Example 59: Exploring similar derivatives


Find the derivatives of
a. F
1 (x) = (1 − x )
2
,
b. F
2 (x) = (1 − x ) ,
3
and
c. F
3 (x)
4
= (1 − x ) .

We'll see later why we are using subscripts for different functions and an uppercase F .
Solution
In order to use the rules we already have, we must first expand each function as
a. F
1 (x) = 1 − 2x + x
2
,
b. F
2 (x) = 1 − 3x + 3 x
2
−x
3
and
c. F
3 (x) = 1 − 4x + 6 x
2
− 4x
3
+x
4
.
It is not hard to see that:

F (x) = −2 + 2x
1

′ 2
F (x) = −3 + 6x − 3x
2

′ 2 3
F (x) = −4 + 12x − 12 x + 4x .
3

An interesting fact is that these can be rewritten as


′ ′ 2 ′ 3
F (x) = −2(1 − x), F (x) = −3(1 − x ) and F (x) = −4(1 − x ) . (2.5.1)
1 2 3

A pattern might jump out at you. Recognize that each of these functions is a composition, letting g(x) = 1 − x :
2
F1 (x) = f1 (g(x)), where f1 (x) = x ,
3
F2 (x) = f2 (g(x)), where f2 (x) = x ,
4
F3 (x) = f3 (g(x)), where f3 (x) = x .

We'll come back to this example after giving the formal statements of the Chain Rule; for now, we are just illustrating a pattern.

Theorem 18: The Chain Rule


Let y = f (u) be a differentiable function of u and let u = g(x) be a differentiable function of x . Then y = f (g(x)) is a
differentiable function of x,and
′ ′ ′
y = f (g(x)) ⋅ g (x). (2.5.2)

To help understand the Chain Rule, we return to Example 59.

2.5.1 https://math.libretexts.org/@go/page/4161
Example 60: Using the Chain Rule
Use the Chain Rule to find the derivatives of the following functions, as given in Example 59.
Solution
Example 59 ended with the recognition that each of the given functions was actually a composition of functions. To avoid
confusion, we ignore most of the subscripts here.
F1 (x) = (1 − x )
2
:
We found that
2 2
y = (1 − x ) = f (g(x)), where f (x) = x and g(x) = 1 − x. (2.5.3)

To find y , we apply the Chain Rule. We need f


′ ′
(x) = 2x and g ′
(x) = −1.

Part of the Chain Rule uses f (g(x)). This means substitute



g(x) for x in the equation for f (x)

. That is, ′
f (x) = 2(1 − x) .
Finishing out the Chain Rule we have
′ ′ ′
y = f (g(x)) ⋅ g (x) = 2(1 − x) ⋅ (−1) = −2(1 − x) = 2x − 2. (2.5.4)

F2 (x) = (1 − x )
3
:
Let y = (1 − x ) = f (g(x)) ,where
3
f (x) = x
3
and g(x) = (1 − x) . We have ′
f (x) = 3 x
2
,so ′
f (g(x)) = 3(1 − x )
2
. The
Chain Rule then states
′ ′ ′ 2 2
y = f (g(x)) ⋅ g (x) = 3(1 − x ) ⋅ (−1) = −3(1 − x ) . (2.5.5)

F3 (x) = (1 − x )
4
:
Finally, when y = (1 − x) ,we have f (x) = x and g(x) = (1 − x) . Thus f
4 4 ′
(x) = 4 x
3
and f ′
(g(x)) = 4(1 − x )
3
. Thus
′ ′ ′ 3 3
y = f (g(x)) ⋅ g (x) = 4(1 − x ) ⋅ (−1) = −4(1 − x ) . (2.5.6)

Example 60 demonstrated a particular pattern: when f (x) = x


n
,then y

= n ⋅ (g(x))
n−1 ′
⋅ g (x) . This is called the Generalized
Power Rule.

Theorem 19: Generalized Power Rule

Let g(x) be a differentiable function and let n ≠ 0 be an integer. Then


d n n−1 ′
(g(x ) ) = n ⋅ (g(x)) ⋅ g (x). (2.5.7)
dx

This allows us to quickly find the derivative of functions like y = (3 x


2
− 5x + 7 + sin x )
20
. While it may look intimidating, the
Generalized Power Rule states that
′ 2 19
y = 20(3 x − 5x + 7 + sin x ) ⋅ (6x − 5 + cos x). (2.5.8)

Treat the derivative--taking process step--by--step. In the example just given, first multiply by 20, then rewrite the inside of the
parentheses, raising it all to the 19 power. Then think about the derivative of the expression inside the parentheses, and multiply by
th

that.
We now consider more examples that employ the Chain Rule.

Example 61: Using the Chain Rule


Find the derivatives of the following functions:
a. y = sin 2x
b. y = ln(4x 3
− 2x )
2

c. y = e −x

Solution
a. Consider y = sin 2x. Recognize that this is a composition of functions, where f (x) = sin x and g(x) = 2x . Thus

2.5.2 https://math.libretexts.org/@go/page/4161
′ ′ ′
y = f (g(x)) ⋅ g (x) = cos(2x) ⋅ 2 = 2 cos 2x. (2.5.9)

b. Recognize that y = ln(4x 3


− 2x )
2
is the composition of f (x) = ln x and g(x) = 4x 3
− 2x
2
. Also, recall that
d 1
( ln x) = . (2.5.10)
dx x

This leads us to:


2
1 12 x − 4x 4x(3x − 1) 2(3x − 1)
′ 2
y = ⋅ (12 x − 4x) = = = . (2.5.11)
3 2 3 2 2
4x − 2x 4x − 2x 2x(2 x − x) 2 x2 − x

c. Recognize that y = e is the composition of f (x) = e and g(x) = −x . Remembering that f ,we have
2
−x x 2 ′ x
(x) = e

2 2
′ −x −x
y =e ⋅ (−2x) = (−2x)e . (2.5.12)

Example 62: Using the Chain Rule to find a tangent line


Let f (x) = cos x . Find the equation of the line tangent to the graph of f at x = 1 .
2

Solution
The tangent line goes through the point (1, f (1)) ≈ (1, 0.54) with slope f ′
(1) . To find f ,we need the Chain Rule.

′ 2
f (x) = − sin(x ) ⋅ (2x) = −2x sin x
2
. Evaluated at x =1 ,we have ′
f (1) = −2 sin 1 ≈ −1.68 . Thus the equation of the
tangent line is
y = −1.68(x − 1) + 0.54. (2.5.13)

The tangent line is sketched along with f in Figure 2.17.

Figure 2.17: f (x) = cos x sketched along with its tangent line at x = 1 .
2

The Chain Rule is used often in taking derivatives. Because of this, one can become familiar with the basic process and learn patterns
that facilitate finding derivatives quickly. For instance,

d 1 (anything)

( ln(anything)) = ⋅ (anything ) = . (2.5.14)
dx anything anything

A concrete example of this is


14 x
d 45 x + sin x + e
15 x
( ln(3 x − cos x + e )) = . (2.5.15)
15 x
dx 3x − cos x + e

While the derivative may look intimidating at first, look for the pattern. The denominator is the same as what was inside the natural
log function; the numerator is simply its derivative.
This pattern recognition process can be applied to lots of functions. In general, instead of writing "anything'', we use u as a generic
function of x. We then say

d u
( ln u) = . (2.5.16)
dx u

The following is a short list of how the Chain Rule can be quickly applied to familiar functions.

2.5.3 https://math.libretexts.org/@go/page/4161
Of course, the Chain Rule can be applied in conjunction with any of the other rules we have already learned. We practice this next.

Example 63: Using the Product, Quotient and Chain Rules


Find the derivatives of the following functions.
a. 5
f (x) = x sin 2 x
3

3
5x
b. f (x) = 2
.
−x
e

Solution
a. We must use the Product and Chain Rules. Do not think that you must be able to "see'' the whole answer immediately; rather,
just proceed step--by--step.
′ 5 2 3 4 3 7 3 4 3
f (x) = x (6 x cos 2 x ) + 5 x ( sin 2 x ) = 6 x cos 2 x + 5x sin 2 x . (2.5.17)

b. We must employ the Quotient Rule along with the Chain Rule. Again, proceed step--by--step.
2 2 2
−x 2 3 −x −x 4 2
e (15 x ) − 5 x ((−2x)e ) e (10 x + 15 x )

f (x) = =
2 2 −2x2
(e
−x
) e

2
x 4 2
=e (10 x + 15 x ).

A key to correctly working these problems is to break the problem down into smaller, more manageable pieces. For instance, when
using the Product and Chain Rules together, just consider the first part of the Product Rule at first: f (x)g (x). Just rewrite f (x),then ′

find g (x). Then move on to the f (x)g(x) part. Don't attempt to figure out both parts at once.
′ ′

Likewise, using the Quotient Rule, approach the numerator in two steps and handle the denominator after completing that. Only
simplify afterward.
We can also employ the Chain Rule itself several times, as shown in the next example.

Example 64: Using the Chain Rule multiple times


Find the derivative of y = tan 5
(6 x
3
− 7x) .
Solution
Recognize that we have the g(x) = tan(6x − 7x) function "inside'' the f (x) = x function; that is, we have
3 5

5
y = ( tan(6 x − 7x)) . We begin using the Generalized Power Rule; in this first step, we do not fully compute the derivative.
3

Rather, we are approaching this step--by--step.


′ 3 4 ′
y = 5( tan(6 x − 7x)) ⋅ g (x). (2.5.18)

We now find g ′
(x) . We again need the Chain Rule;
′ 2 3 2
g (x) = sec (6 x − 7x) ⋅ (18 x − 7). (2.5.19)

Combine this with what we found above to give


′ 3 4 2 3 2
y = 5( tan(6 x − 7x)) ⋅ sec (6 x − 7x) ⋅ (18 x − 7)

2 2 3 4 3
= (90 x − 35) sec (6 x − 7x) tan (6 x − 7x).

This function is frankly a ridiculous function, possessing no real practical value. It is very difficult to graph, as the tangent
function has many vertical asymptotes and 6x − 7x grows so very fast. The important thing to learn from this is that the
3

2.5.4 https://math.libretexts.org/@go/page/4161
derivative can be found. In fact, it is not "hard;'' one must take several simple steps and be careful to keep track of how to apply
each of these steps.

It is a traditional mathematical exercise to find the derivatives of arbitrarily complicated functions just to demonstrate that it can be
done. Just break everything down into smaller pieces.

Example 65: Using the Product, Quotient and Chain Rules


−2 2 4x
x cos(x ) − sin (e )
Find the derivative of f (x) = 2 4
.
ln(x + 5x )

Solution
This function likely has no practical use outside of demonstrating derivative skills. The answer is given below without
simplification. It employs the Quotient Rule, the Product Rule, and the Chain Rule three times.
2 4 −2 −3 −2 4x 4x 4x
( ln(x + 5 x )) ⋅ [(x ⋅ (− sin(x )) ⋅ (−2 x ) + 1 ⋅ cos(x )) − 2 sin(e ) ⋅ cos(e ) ⋅ (4 e )]

3
2x + 20x
−2 2 4x
− (x cos(x ) − sin (e )) ⋅
2 4
′ x + 5x
f (x) = . (2.5.20)
2
( ln(x2 + 5 x4 ))

The reader is highly encouraged to look at each term and recognize why it is there. (I.e., the Quotient Rule is used; in the
numerator, identify the "LOdHI'' term, etc.) This example demonstrates that derivatives can be computed systematically, no
matter how arbitrarily complicated the function is.

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.

Example 66: The Chain Rule and exponential functions


Use the Chain Rule to find the derivative of y = a where a > 0 ,a ≠ 1 is constant.
x

Solution
We only know how to find the derivative of one exponential function: y = e ; this problem is asking us to find the derivative of
x

functions such as y = 2 . x

This can be accomplished by rewriting a in terms of e . Recalling that e and ln x are inverse functions, we can write
x x

x
ln a x ln( a )
a =e and so y =a =e .

By the exponent property of logarithms, we can "bring down'' the power to get
x x(ln a)
y =a =e .

The function is now the composition y = f (g(x)) ,with f (x) = e and g(x) = x(ln a) . Since f x ′
(x) = e
x
and g ′
(x) = ln a , the
Chain Rule gives
′ x(ln a)
y =e ⋅ ln a.

Recall that the e x(ln a)


term on the right hand side is just x
a ,our original function. Thus, the derivative contains the original
function itself. We have
′ x
y = y ⋅ ln a = a ⋅ ln a.

The Chain Rule, coupled with the derivative rule of e ,allows us to find the derivatives of all exponential functions.
x

The previous example produced a result worthy of its own "box.''

2.5.5 https://math.libretexts.org/@go/page/4161
Theorem 20: Derivatives of Exponential Functions
Let f (x) = a ,for a > 0, a ≠ 1 . Then f is differentiable for all real numbers and
x

′ x
f (x) = ln a ⋅ a .

Alternate Chain Rule Notation


dy
It is instructive to understand what the Chain Rule "looks like'' using " '' notation instead of y notation. Suppose that y = f (u) is

dx
a function of u,where u = g(x) is a function of x,as stated in Theorem 18. Then, through the composition f ∘ g ,we can think of y as
dy
a function of x,as y = f (g(x)) . Thus the derivative of y with respect to x makes sense; we can talk about . This leads to an
dx
interesting progression of notation:
′ ′ ′
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
= . (2.5.21)
dx dx

dy
It is important to realize that we are not canceling these terms; the derivative notation of is one symbol. It is equally important to
dx
realize that this notation was chosen precisely because of this behavior. It makes applying the Chain Rule easy with multiple
variables. For instance,
dy dy d◯ d△
= ⋅ ⋅ . (2.5.22)
dt d◯ d△ dt

where ◯ and △ are any variables you'd like to use.


One of the most common ways of "visualizing" the Chain Rule is to consider a set of gears, as shown in Figure 2.18. 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. Using the terminology of calculus,
du
the rate of u-change, with respect to x,is =2 .
dx

dy dy du
Figure 2.18: A series of gears to demonstrate the Chain Rule. Note how = ⋅ .
dx du dx

dy
Likewise, every revolution of u causes 3 revolutions of y : =3 . How does y change with respect to x? For each revolution of x,y
du
revolves 6 times; that is,

2.5.6 https://math.libretexts.org/@go/page/4161
dy dy du
= ⋅ = 2 ⋅ 3 = 6. (2.5.23)
dx du dx

We can then extend the Chain Rule with more variables by adding more gears to the picture.
It is difficult to overstate the importance of the Chain Rule. So often the functions that we deal with are compositions of two or more
functions, requiring us to use this rule to compute derivatives. It is often used in practice when actual functions are unknown. Rather,
dy du dy
through measurement, we can calculate and . With our knowledge of the Chain Rule, finding is straightforward.
du dx dx

In the next section, we use the Chain Rule to justify another differentiation technique. There are many curves that we can draw in the
plane that fail the "vertical line test.'' For instance, consider x + y = 1 ,which describes the unit circle. We may still be interested in
2 2

dy
finding slopes of tangent lines to the circle at various points. The next section shows how we can find without first "solving for
dx
y .'' While we can in this instance, in many other instances solving for y is impossible. In these situations, implicit differentiation is
indispensable.

This page titled 2.5: The Chain Rule is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et al.
via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

2.5.7 https://math.libretexts.org/@go/page/4161
2.6: Implicit Differentiation
dy
In the previous sections we learned to find the derivative, , or y , when y is given explicitly as a function of x. That is, if we
dx

know y = f (x) for some function f , we can find y . For example, given y = 3x − 7 , we can easily find y = 6x . (Here we
′ 2 ′

explicitly state how x and y are related. Knowing x, we can directly find y .)
Sometimes the relationship between y and x is not explicit; rather, it is implicit. For instance, we might know that x − y = 4 . 2

This equality defines a relationship between x and y ; if we know x, we could figure out y . Can we still find y ? In this case, sure; ′

we solve for y to get y = x − 4 (hence we now know y explicitly) and then differentiate to get y = 2x .
2 ′

Sometimes the implicit relationship between x and y is complicated. Suppose we are given sin(y) + y = 6 − x . A graph of this 3 3

implicit function is given in Figure 2.19. In this case there is absolutely no way to solve for y in terms of elementary functions. The
surprising thing is, however, that we can still find y via a process known as implicit differentiation.

Figure 2.19: A graph of the implicit function sin(y) + y 3


= 6−x
2
.
Implicit differentiation is a technique based on 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).
We begin by reviewing the Chain Rule. Let f and g be functions of x. Then
d
′ ′
(f (g(x))) = f (g(x)) ⋅ g (x). (2.6.1)
dx

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


d ′ ′
d ′
dy
(f (y))) = f (y)) ⋅ y , or (f (y))) = f (y) ⋅ . (2.1)
dx dx dx

These equations look strange; the key concept to learn here is that we can find y even if we don't exactly know how y and x relate.

We demonstrate this process in the following example.

Example 67: Using Implicit Differentiation


Find y given that sin(y) + y
′ 3
= 6 −x
3
.
Solution
We start by taking the derivative of both sides (thus maintaining the equality.) We have :
d 3
d 3
( sin(y) + y ) = (6 − x ). (2.6.2)
dx dx

The right hand side is easy; it returns −3x . 2

The left hand side requires more consideration. We take the derivative term--by--term. Using the technique derived from
Equation 2.1 above, we can see that
d ′
( sin y) = cos y ⋅ y . (2.6.3)
dx

We apply the same process to the y term. 3

2.6.1 https://math.libretexts.org/@go/page/4162
d 3
d 3 2 ′
(y ) = ((y ) ) = 3(y ) ⋅y . (2.6.4)
dx dx

Putting this together with the right hand side, we have


′ 2 ′ 2
cos(y)y + 3 y y = −3 x . (2.6.5)

Now solve for y .′

′ 2 ′ 2
cos(y)y + 3 y y = −3 x .

2 ′ 2
( cos y + 3 y )y = −3x

2

−3x
y =
2
cos y + 3y

This equation for y probably seems unusual for it contains both x and y terms. How is it to be used? We'll address that next.

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 computing 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 easier 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 (√6, 0) lies on the graph of the implicit function sin y + y = 6 − x . Plugging in
3 3 3


0 for y , we see the left hand side is 0 . Setting x = √6 , we see the right hand side is also 0 ; the equation is satisfied. The following
3

example finds the equation of the tangent line to this function at this point.

Example 68: Using Implicit Differentiation to find a tangent line



Find the equation of the line tangent to the curve of the implicitly defined function sin y + y at the point (√6, 0).
3
3 3
= 6 −x

Solution
In Example 67 we found that
2
−3x

y = . (2.6.6)
2
cos y + 3y

– 3 – 3 –

We find the slope of the tangent line at the point (√6, 0) by substituting for x and 0 for y . Thus at the point , we
3
√6 (√6, 0)

have the slope as


3 – −−
2 3
−3(√6) −3 √36

y = = ≈ −9.91. (2.6.7)
2
cos 0 + 3 ⋅ 0 1


Therefore the equation of the tangent line to the implicitly defined function sin y + y 3
= 6 −x
3
at the point (√6, 0) is
3

3 −− 3 –
y = −3 √36(x − √6) + 0 ≈ −9.91x + 18. (2.6.8)

The curve and this tangent line are shown in Figure 2.20.


Figure 2.20: The function sin y + y 3 2
and its tangent line at the point (√6, 0).
3
= 6−x

This suggests a general method for implicit differentiation. For the steps below assume y is a function of x.

2.6.2 https://math.libretexts.org/@go/page/4162
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 except that, because of the Chain Rule, 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.

3. Factor out y ; solve for y by dividing.


′ ′

dy
Practical Note: When working by hand, it may be beneficial to use the symbol dx
instead of y

, as the latter can be easily
confused for y or y . 1

Example 69: Using Implicit Differentiation


Given the implicitly defined function y 3
+x y
2 4
= 1 + 2x , find y . ′

Solution
We will take the implicit derivatives term by term. The derivative of y is 3y 3 2
y

.
The second term, x y , is a little tricky. It requires the Product Rule as it is the product of two functions of x: x and y . Its
2 4 2 4

derivative is x (4y y ) + 2x y . The first part of this expression requires a y because we are taking the derivative of a y term.
2 3 ′ 4 ′

The second part does not require it because we are taking the derivative of x . 2

The derivative of the right hand side is easily found to be 2. In all, we get:
2 ′ 2 3 ′ 4
3 y y + 4 x y y + 2x y = 2. (2.6.9)

Move terms around so that the left side consists only of the y terms and the right side consists of all the other terms:

2 ′ 2 3 ′ 4
3y y + 4x y y = 2 − 2x y . (2.6.10)

Factor out y from the left side and solve to get


4
2 − 2xy

y = . (2.6.11)
2 2 3
3y + 4x y

To confirm the validity of our work, let's find the equation of a tangent line to this function at a point. It is easy to confirm that
the point (0, 1) lies on the graph of this function. At this point, y = 2/3 . So the equation of the tangent line is ′

y = 2/3(x − 0) + 1 . The function and its tangent line are graphed in Figure 2.21.

Figure 2.21: A graph of the implicitly defined function y 3


+x y
2 4
= 1 + 2x along with its tangent line at the point (0, 1).
Notice how our function looks much different than other functions we have seen. For one, it fails the vertical line test. Such
functions are important in many areas of mathematics, so developing tools to deal with them is also important.

Example 70: Using Implicit Differentiation

Given the implicitly defined function sin(x 2 2


y )+y
3
= x +y , find y . ′

Solution
Differentiating term by term, we find the most difficulty in the first term. It requires both the Chain and Product Rules.
d d
2 2 2 2 2 2
( sin(x y )) = cos(x y ) ⋅ (x y )
dx dx
2 2 2 ′ 2
= cos(x y ) ⋅ (x (2y y ) + 2x y )

2 ′ 2 2 2
= 2(x y y + x y ) cos(x y ).

2.6.3 https://math.libretexts.org/@go/page/4162
We leave the derivatives of the other terms to the reader. After taking the derivatives of both sides, we have
2 ′ 2 2 2 2 ′ ′
2(x y y + x y ) cos(x y ) + 3 y y = 1 +y . (2.6.12)

We now have to be careful to properly solve for y , particularly because of the product on the left. It is best to multiply out the

product. Doing this, we get


2 2 2 ′ 2 2 2 2 ′ ′
2 x y cos(x y )y + 2x y cos(x y ) + 3 y y = 1 +y . (2.6.13)

From here we can safely move around terms to get the following:
2 2 2 ′ 2 ′ ′ 2 2 2
2 x y cos(x y )y + 3 y y − y = 1 − 2x y cos(x y ). (2.6.14)

Then we can solve for y to get


2 2 2
1 − 2x y cos(x y )

y = . (2.6.15)
2 2 2 2
2 x y cos(x y ) + 3 y −1

A graph of this implicit function is given in Figure 2.22. 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 . ′

Figure 2.22: A graph of the implicitly defined function sin(x 2 2


y ) +y
3
= x+y .
At (0, 0), the slope is −1.
At (0, 1), the slope is 1/2.
At (0, −1), the slope is also 1/2.
The tangent lines have been added to the graph of the function in Figure 2.23.

Figure 2.23: A graph of the implicitly defined function sin(x 2 2


y ) +y
3
= x+y and certain tangent lines.

Quite a few "famous'' curves have equations that are given implicitly. We can use implicit differentiation to find the slope at various
points on those curves. We investigate two such curves in the next examples.

Example 71: Finding slopes of tangent lines to a circle



Find the slope of the tangent line to the circle x 2
+y
2
=1 at the point (1/2, √3/2).
Solution
Taking derivatives, we get 2x + 2y y ′
=0 . Solving for y gives: ′

2.6.4 https://math.libretexts.org/@go/page/4162
−x

y = . (2.6.16)
y

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

At the point (1/2, √3/2), we have the tangent line's slope as
−1/2 −1

y = – = – ≈ −0.577. (2.6.17)
√3/2 √3


A graph of the circle and its tangent line at (1/2, √3/2) is given in Figure 2.24, along with a thin dashed line from the origin
that is perpendicular to the tangent line. (It turns out that all normal lines to a circle pass through the center of the circle.)


Figure 2.24: The unit circle with its tangent line at (1/2, √3/2).

This section has shown how to find the derivatives of implicitly defined functions, whose graphs include a wide variety of
interesting and unusual shapes. Implicit differentiation can also be used to further our understanding of "regular'' differentiation.
One hole in our current understanding of derivatives is this: what is the derivative of the square root function? That is,
d − d 1/2
(√x ) = (x ) =? (2.6.18)
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 and obtain
d 1/2
1 −1/2
1
(x ) = x = −. (2.6.19)
dx 2 2 √x

The trouble with this is that the Power Rule was initially defined only for positive integer powers, n > 0 . While we did not justify
this at the time, generally the Power Rule is proved using something called the Binomial Theorem, which deals only with positive
integers. The Quotient Rule allowed us to extend the Power Rule to negative integer powers. Implicit Differentiation allows us to
extend the Power Rule to rational powers, as shown below.
Let y = x m/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 = x . Now apply implicit differentiation.
n m

2.6.5 https://math.libretexts.org/@go/page/4162
m/n
y =x
n m
y =x

d n
d m
(y ) = (x )
dx dx
n−1 ′ m−1
n⋅y ⋅y =m⋅x
m−1
m x
′ m/n
y = (now substitute x for y)
n−1
n y
m−1
m x
= (apply lots of algebra)
n (xm/n )n−1

m (m−n)/n
= x
n
m
m/n−1
= x .
n

The above derivation is the key to the proof extending the Power Rule to rational powers. Using limits, we can extend this once
more to include all powers, including irrational (even transcendental!) powers, giving the following theorem.

Theorem 21: Power Rule for Differentiation

Let f (x) = x , where n ≠ 0 is a real number. Then f is a differentiable function, and f


n ′ n−1
(x) = n ⋅ x .

This theorem allows us to say the derivative of x is πx π π−1


.
We now apply this final version of the Power Rule in the next example, the second investigation of a "famous'' curve.

Example 72: Using the Power Rule

Find the slope of x 2/3


+y
2/3
=8 at the point (8, 8).
Solution
This is a particularly interesting curve called an astroid. It is the shape traced out by a point on the edge of a circle that is
rolling around inside of a larger circle, as shown in Figure 2.25.

Figure 2.25: An astroid, traced out by a point on the smaller circle as it rolls inside the larger circle.
To find the slope of the astroid at the point (8, 8), we take the derivative implicitly.
2 2
−1/3 −1/3 ′
x + y y =0
3 3

2 −1/3 ′
2 −1/3
y y =− x
3 3
−1/3

x
y =−
y −1/3

1/3 −

y y

y =− = −√
3
.
1/3 x
x

Plugging in x = 8 and y = 8 , we get a slope of −1. The astroid, with its tangent line at (8, 8), is shown in Figure 2.26.

2.6.6 https://math.libretexts.org/@go/page/4162
Figure 2.26: An astroid with a tangent line.

Implicit Differentiation and the Second Derivative


dy
We can use implicit differentiation to find higher order derivatives. In theory, this is simple: first find , then take its derivative dx

with respect to x. In practice, it is not hard, but it often requires a bit of algebra. We demonstrate this in an example.

Example 73: Finding the second derivative


2
d y
Given x 2
+y
2
=1 , find dx
2
=y
′′
.
Solution
dy
We found that y ′
=
dx
= −x/y in Example 71. To find y , we apply implicit differentiation to y .
′′ ′

d
′′ ′
y = (y )
dx

d x
= (− ) (Now use the Quotient Rule.)
dx y

y(1) − x(y )
=−
2
y

replace y with −x/y:


y − x(−x/y)
=−
y2
2
y + x /y
=− .
2
y

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.

Logarithmic Differentiation
Consider the function y = x ; it is graphed in Figure 2.27. It is well--defined for
x
x >0 and we might be interested in finding
equations of lines tangent and normal to its graph. How do we take its derivative?

2.6.7 https://math.libretexts.org/@go/page/4162
Figure 2.27: A plot of y = x .
x

The function is not a power function: it has a "power'' of x, not a constant. It is not an exponential function: it has a "base'' of x, not
a constant.
A differentiation technique known as logarithmic differentiation becomes useful here. The basic principle is this: take the natural
log of both sides of an equation y = f (x), then use implicit differentiation to find y . We demonstrate this in the following

example.

Example 74: Using Logarithmic Differentiation


Given y = x , use logarithmic differentiation to find y .
x ′

Solution
As suggested above, we start by taking the natural log of both sides then applying implicit differentiation.
x
y =x
x
ln(y) = ln(x )(apply logarithm rule)

ln(y) = x ln x(now use implicit differentiation)

d d
( ln(y)) = (x ln x)
dx dx

y 1
= ln x + x ⋅
y x

y
= ln x + 1
y
′ x
y = y( ln x + 1)(substitute y = x )

′ x
y = x ( ln x + 1).

To "test'' our answer, let's use it to find the equation of the tangent line at x = 1.5. The point on the graph our tangent line must
pass through is (1.5, 1.5 ) ≈ (1.5, 1.837). Using the equation for y , we find the slope as
1.5 ′

′ 1.5
y = 1.5 ( ln 1.5 + 1) ≈ 1.837(1.405) ≈ 2.582. (2.6.20)

Thus the equation of the tangent line is y = 1.6833(x − 1.5) + 1.837. Figure 2.28 graphs y = x along with this tangent line.
x

Figure 2.22: A graph of y = x and its tangent line at x = 1.5 .


x

2.6.8 https://math.libretexts.org/@go/page/4162
Implicit differentiation proves to be useful as it allows us to find the instantaneous rates of change of a variety of functions. In
particular, it extended the Power Rule to rational exponents, which we then extended to all real numbers. In the next section,
implicit differentiation will be used to find the derivatives of inverse functions, such as y = sin x . −1

This page titled 2.6: Implicit Differentiation is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory
Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

2.6.9 https://math.libretexts.org/@go/page/4162
2.7: Derivatives of Inverse Functions
Recall that a function y = f (x) is said to be one to one if it passes the horizontal line test; that is, for two different x values x and 1

x , we do not have f (x ) = f (x ) . In some cases the domain of f must be restricted so that it is one to one. For instance, consider
2 1 2

f (x) = x . Clearly, f (−1) = f (1), so f is not one to one on its regular domain, but by restricting f to (0, ∞), f is one to one.
2

Now recall that one to one functions have inverses. That is, if f is one to one, it has an inverse function, denoted by f −1
, such that
if f (a) = b , then f (b) = a . The domain of f is the range of f , and vice-versa. For ease of notation, we set g = f
−1 −1 −1
and treat
g as a function of x.

Since f (a) = b implies g(b) = a , when we compose f and g we get a nice result:

f (g(b)) = f (a) = b. (2.7.1)

In general, f (g(x)) = x and g(f (x)) = x . This gives us a convenient way to check if two functions are inverses of each other:
compose them and if the result is x, then they are inverses (on the appropriate domains.)
When the point (a, b) lies on the graph of f , the point (b, a) lies on the graph of g . This leads us to discover that the graph of g is
the reflection of f across the line y = x . In Figure 2.29 we see a function graphed along with its inverse. See how the point (1, 1.5)
lies on one graph, whereas (1.5, 1) lies on the other. Because of this relationship, whatever we know about f can quickly be
transferred into knowledge about g .

Figure 2.29: A function f along with its inverse f


−1
. (Note how it does not matter which function we refer to as f ; the other is
f .)
−1

For example, consider Figure 2.30 where the tangent line to f at the point (a, b) is drawn. That line has slope f (a) . Through ′

reflection across y = x , we can see that the tangent line to g at the point (b, a) should have slope . This then tells us that

f (a)
1

′ 1
g (b) = ′
.
f (a)

Figure 2.30: Corresponding tangent lines drawn to f and f .


−1

Consider:

2.7.1 https://math.libretexts.org/@go/page/4636
We have discovered a relationship between f and g in a mostly graphical way. We can realize this relationship analytically as
′ ′

well. Let y = g(x) , where again g = f . We want to find y . Since y = g(x) , we know that f (y) = x . Using the Chain Rule and
−1 ′

Implicit Differentiation, take the derivative of both sides of this last equality.
d d
(f (y)) = (x)
dx dx
′ ′
f (y) ⋅ y =1


1
y =

f (y)


1
y =

f (g(x))

This leads us to the following theorem.

Theorem 22: Derivatives of Inverse Functions


Let f be differentiable and one to one on an open interval I , where f (x) ≠ 0 for all x in I , let J be the range of f on I , let g

be the inverse function of f , and let f (a) = b for some a in I . Then g is a differentiable function on J , and in particular,

1. (f
−1 ′
) (b) = g (b) = ′
1
\quad \text{ and }\quad 2. \left(f^{-1}\right)^\prime (x)=g^\prime (x) = \frac{1}
f (a)

{f^\prime(g(x))}\)

The results of Theorem 22 are not trivial; the notation may seem confusing at first. Careful consideration, along with examples,
should earn understanding.
In the next example we apply Theorem 22 to the arcsine function.

Example 75: Finding the derivative of an inverse trigonometric function

Let y = arcsin x = sin −1


x . Find y using Theorem 22.

Figure 2.32: A right triangle defined by y = sin −1


(x/1) with the length of the third leg found using the Pythagorean Theorem.
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 2.32. Recall that the sine
function can 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

2.7.2 https://math.libretexts.org/@go/page/4636
y = arcsin x can be rewritten as y = arcsin(x/1); that is, consider a right triangle where the hypotenuse has length 1 and the
−−−− −
side opposite of the angle with measure y has length x. This means the final side has length √1 − x , using the Pythagorean
2

Theorem.
− −−− − − −−− −
−1 2 2
Therefore cos(sin x) = cos y = √ 1 − x /1 = √ 1 − x , resulting in (2.7.2)

d 1

( arcsin x) = g (x) = − −−− −. (2.7.3)
dx √ 1 − x2

Remember that the input x of the arcsine function is a ratio of a side of a right triangle to its hypotenuse; the absolute value of this
ratio will never be greater than 1. Therefore the inside of the square root will never be negative.
In order to make y = sin x one to one, we restrict its domain to [−π/2, π/2]; on this domain, the range is [−1, 1]. Therefore the
domain of y = arcsin x is [−1, 1] and the range is [−π/2, π/2]. When x = ±1 , note how the derivative of the arcsine function is
undefined; this corresponds to the fact that as x → ±1 , the tangent lines to arcsine approach vertical lines with undefined slopes.
In Figure 2.33 we see f (x) = sin x and f = sin x graphed on their respective domains. The line tangent to sin x at the point
−1 −1

– –
(π/3, √3/2) has slope cos π/3 = 1/2. The slope of the corresponding point on sin x, the point (√3/2, π/3), is
−1

1 1 1 1
− −−− − −−−−− = − −−−− − = − −− = = 2, (2.7.4)
– √ 1 − 3/4 √1/4 1/2
√ 1 − (√3/2 )2

verifying yet again that at corresponding points, a function and its inverse have reciprocal slopes.

Figure 2.33: Graphs of sin x and sin


−1
x along with corresponding tangent lines.
Using similar techniques, we can find the derivatives of all the inverse trigonometric functions. In Figure 2.31 we show the
restrictions of the domains of the standard trigonometric functions that allow them to be invertible.

2.7.3 https://math.libretexts.org/@go/page/4636
Figure 2.31: Domains and ranges of the trigonometric and inverse trigonometric functions.

Theorem 23: Derivatives of Inverse Trigonometric Functions

The inverse trigonometric functions are differentiable on all open sets contained in their domains (as listed in Figure 2.31) and
their derivatives are as follows:
d −1
1 d −1
1
1. ( sin (x)) = − −−− − 4. ( cos (x)) = − − −−− − (2.7.5)
dx √ 1 − x2 dx √ 1 − x2

d 1 d 1
−1 −1
2. ( sec (x)) = 5. ( csc (x)) = − (2.7.6)
− −−−− − −−−−
dx 2
|x| √ x − 1 dx 2
|x| √ x − 1

d −1
1 d −1
1
3. ( tan (x)) = 6. ( cot (x)) = − (2.7.7)
2 2
dx 1 +x dx 1 +x

Note how the last three derivatives are merely the opposites of the first three, respectively. Because of this, the first three are used
almost exclusively throughout this text.
In Section 2.3, we stated without proof or explanation that d

dx
( ln x) =
1

x
. We can justify that now using Theorem 22, as shown in
the example.

Example 76: Finding the derivative of y = ln x


Use Theorem 22 to compute d

dx
( ln x) .
Solution
View y = ln x as the inverse of y =e
x
. Therefore, using our standard notation, let f (x) = e
x
and g(x) = ln x . We wish to
find g (x). Theorem 22 gives:

1

g (x) =

f (g(x))

1
=
eln x
1
= .
x

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.

Theorem 24: Glossary of Derivatives of Elementary Functions


Let u and v be differentiable functions, and let a , c and n be real numbers, a > 0 , n ≠ 0 .

2.7.4 https://math.libretexts.org/@go/page/4636
This page titled 2.7: Derivatives of Inverse Functions is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by
Gregory Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available
upon request.

2.7.5 https://math.libretexts.org/@go/page/4636
2.E: Applications of Derivatives(Exercises)
2.1: Instantaneous Rates of Change: The Derivative
Terms and Concepts
1. T/F: 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. T/F: 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 7 and 10.
5. Let y = f (x). Give three different notations equivalent to “f '(x). ”

Problems
In Exercises 6-12, use the definition of the derivative to compute the derivative of the given function.
6. f (x) = 6
7. f (x) = 2x
8. f (t) = 4 − 3t
9. g(x) = x 2

10. f (x) = 3x 2
−x +4

11. r(x) = x
1

12. r(s) = s−2


1

In Exercises 13-19, a function and an x-value c are given.


(Note: these functions are the same as those given in Exercises 6 through 12.)
(a) Find the tangent line to the graph of the function at c .
(b) Find the normal line to the graph of the function at c .
13. f (x) = 6, at x = −2 .
14. f (x) = 2x, at x = 3 .
15. f (x) = 4 − 3x, at x = 7 .
16. g(x) = x 2
, at x = 2 .
17. f (x) = 3x 2
− x + 4, at x = −1 .
18. r(x) = 1

x
, at x = −2 .
19. r(x) = x−2
1
, at x = 3 .
In Exercises 20-23, 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 . ′

20. f (x) = x 2
+ 2x + 1, x = 3

21. f (x) = x+1


10
, x =9

22. f (x) = e x
, x =2

23. f (x) = cos x, x =0

24. The graph of f (x) = x − 1 is shown. 2

(a) Use the graph to approximate the slope of the tangent line to f at the following points: (-1,0),(0,-1) and (2,3).
(b) Using the definition, find f (x). ′

2.E.1 https://math.libretexts.org/@go/page/9969
(c) Find the slope of the tangent line at the points (-1,0), (0,-1) and (2,3).

25. The graph of f (x) = 1

x+1
is shown.
(a) Use the graph to approximate the slope of the tangent line to f at the following points: (0,1) and (1, 0.5).
(b) Using the definition, find f (x).

(c) Find the slope of the tangent line at the points (0, 1) and (1, 0.5).

In Exercises 26-29, a graph of a function f (x) is given. Using the graph, sketch f ′
.
(x)

26.

27.

28.

2.E.2 https://math.libretexts.org/@go/page/9969
29.

30. Using the graph of g(x) below, answer the following questions.
(a) Where is g(x) > 0 ?
(b) Where is g(x) < 0 ?
(c) Where is g(x) = 0 ?
(d) Where is g (x) < 0 ?

(e) Where is g (x) > 0 ?


(f) Where is g (x) = 0 ?


Review
2

31. Approximate lim x +2x−35

x2 −10.5+27.5
.
x→5

32. Use the Bisection Method to approximate, accurate to two decimal places, the root of g(x) = x
3
+x
2
+x −1 on [0.5, 0.6].
33. Give intervals on which each of the following functions are continuous.
(a) 1

ex +1

(b) x
1

e −1
−−−−−
(c) √5 − x
−−−− −
(d) √5 − x 2

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


(a) lim f (x) =?

x→−3

(b) lim
+
f (x) =?
x→−3

(c) lim f (x) =?


x→−3

(d) Where is f continuous?

2.2: Interpretations of the Derivative


Terms and Concepts
1. What is the instantaneous rate of change of position called?
2. Given a function y = f (x), in your own words describe how to find the units of f '(x).
3. What functions have a constant rate of change?

2.E.3 https://math.libretexts.org/@go/page/9969
Problems
4. Given f (5 = 10 and f ′
(5) = 2 , approximate f (6).
5. Given P (100) = −67 and P '(100) = 5, approximate P (110).
6. Given z(25) = 187 and z'(25) = 17 , approximate z(20).
7. Knowing f (10) = 25 and f '(10) = 5 and the methods described in this section, which approximation is likely to be most
accurate: f (10.1), f (11), or f (20)? Explain your reasoning.
8. Given f (7) = 26 and f (8) = 22 , approximate f '(7).
9. Given H (0) = 17 and H (2) = 29 , approximate H '(2).
10. Let V (x) measure the volume, in decibels, measured inside a restaurant with x customers. What are the units of V '(x)?
11. Let v(t) measure the velocity, in ft/s, of a car moving in a straight line t seconds after starting. What are the units of v'(t) ?
12. The height H, in feet, of a river is recorded t hours after midnight, April 1. What are the units of H '(t)?
13. P is the profit, in thousands of dollars, of producing and selling c cars.
(a) What are the units of P (c)?

(b) What is likely true of P (0P ?


14. T is the temperature in degrees Fahrenheit, h hours after midnight on July 4 in Sidney, NE.
(a) What are the units of T '(h)?
(b) Is T '(8) likely greater than or less than 0? Why?
(c) Is T (8) likely greater than or less than 0? Why?
In Exercises 15-18, graphs of the functions f (x) and g(x) are given. Identify which function is the derivative of the other.
15.

16.

2.E.4 https://math.libretexts.org/@go/page/9969
17.

18.

Review
In Exercises 19-20, use the definition to compute the derivatives of the following functions.
19. f (x) = 5x 2

20. f (x) = √−
x at x = 9 .

In Exercises 21-22, numerically approximate the value of f ′


(x) at the indicated x value.
21. f (x) = cos x at x = π .
22. f (x) = √−
x at x = 9 .

2.3: Basic Differentiation Rules


Terms and Concepts
1. What is the name of the rule which states that d

dx
n n−1
(x ) = nx , where n > 0 is an integer?

2. What is d

dx
(ln x) ?
3. Give an example of a function f(x) where f '(x) = f (x).
4. Give an example of a function f (x) where f '(x) = 0.
5. The derivative rules introduced in this section explain how to compute the derivative of which of the following functions?
3
f (x) =
x2
2
g(x) = 3 x − x + 17

h(x) = 5 ln x

j(x) = sin x cos x


2
x
k(x) = e

m(x) = √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.”

2.E.5 https://math.libretexts.org/@go/page/9969
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, where x is measured in feet. What are the units of f ''(x)?

Problems
In Exercises 11-25, compute the derivative of the given function.
11. f (x) = 7x 2
− 5x + 7

12. g(x) = 14x 3


+ 7x
2
+ 11x − 29

13. m(t) = 9t 2

1

8
3
t + 3t − 8

14. f (θ) = 9 sin θ + 10 cos θ


15. f (r) = 6e r

16. g(t) = 10t 4


− cos t + 7 sin t

17. f (x) = 2 ln x − x
18. p(s) = 1

4
s
4
+
1

3
s
3
+
1

2
2
s +s+1

19. h(t) = e t
− sin t − cos t

20. f (x) = ln(5x 2


)

21. f (t) = ln(17) + e 2


+ sin π/2

22. g(t) = (1 + 3t) 2

23. g(x) = (2x − 5) 3

24. f (x) = (1 − x) 3

25. f (x) = (2 − 3x) 2

logb x
26. A property of logarithms is that log a
x =
log a
, for all bases a, b>0, ≠ 1 .
b

(a) Rewrite this identity when b = e , i.e., using log ex = ln x.


(b) Use part (a) to find the derivative of y = log x . a

(c) Give the derivative of y = log x . 10

In Exercises 27-32, compute the first four derivatives of the given function.
27. f (x) = x 6

28. g(x) = 2 cos x


29. h(t) = t 2
−e
t

30. p(θ) = θ 4
−θ
3

31. f (θ) = sin θ − cos θ


32. f (x) = 1, 100
In Exercises 33-38, find the equations of the tangent and normal lines to the graph of the function at the given point.
33. f (x) = x 3
− x at x = 1

34. f (t) = e t
at t = 0

35. g(x) = ln x at t = 0
36. f (x) = 4 sin x at x = π/2
37. f (x) = −2 cos x at x = π/4
38. f (x) = 2x + 3 at x = 5

2.E.6 https://math.libretexts.org/@go/page/9969
Review
39. Given that e 0
=1 , approximate the value of e 0.1
using the tangent line to f (x) = e x
at x = 0 .
40. Approximate the value of (3.01) using the tangent line to f (x) = x
4 4
at x = 3 .

2.4: The Product and Quotient Rules


Terms and Concepts
1. T/F: The Product Rule states that f racddx (x 2
sin x) = 2x cos x .
2

2. T/F: The Quotient Rule states that dx


d
(
x

sin x
) =
cos x

2x
.

3. T/F: 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. T/F: Regardless of the function, there is always exactly one right way of computing its derivative.
6. In your own words, explain what it means to make your answers "clear."

Problems
In Exercises 7-10:
(a) Use the Product Rule to differentiate the function.
(b) Manipulate the function algebraically and differentiate without the Product Rule.
(c) Show that the answers from (a) and (b) are equivalent.
7. f (x) = x(x 2
+ 3x)

8. g(x) = 2x 2
(5 x )
3

9. h(s) = (2s − 1)(s + 4)


10. f (x) = (x 2
+ 5)(3 − x )
3

In Exercises 11-14:
(a) Use the Quotient Rule to differentiate the function.
(b) Manipulate the function algebraically and differentiate without the Quotient Rule.
(c) Shows that the answers from (a) and (b) are equivalent.
2
x +3
11. f (x) = x

3 2

12. g(x) = x −2 x

2x
2

13. h(s) = 4s
3
3

14. f (t) = t −1

t+1

In Exercises 15-29, compute the derivative of the given function.


15. f (x) = x sin x
16. f (t) = 1
2
(csc t − 4)
t

x+7
17. g(x) = x−5

18. g(t) = cos t−2t


t
2

19. h(x) = cot x − e x

20. h(t) = 7t 2
+ 6t − 2

4 3

21. f (x) = x +2 x

x
2

22. f (x) = (16x 3


+ 24 x
2
+ 3x)
7x−1

16 x3 +24 x2 +3x

23. f (t) = t 5
(sec t + e )
t

2.E.7 https://math.libretexts.org/@go/page/9969
24. f (x) = cos x+3
sin x

25. g(x) = e 2
(sin(π/4) − 1)

26. g(t) = 4t 3 t
e − sin t cos t

27. h(t) = t
2
sin t+3

t cos t+2

28. f (x) = x 2
e
x
tan x

29. g(x) = 2x sin x sec x


In Exercises 30-33, find the equations of the tangent and normal lines to the graph of g at the indicated point.
30. g(s) = e s
(s
2
+ 2) at (0, 2) .

31. g(t) = t sin t at (


2
,
−3π

2
)

32. g(x) = x−1


x
at (2, 4) .

33. g(θ) = cos θ−8θ

θ+1
at (0, −5)

In Exercises 34-37, find the x-values where the graph of the function has a horizontal tangent line.
34. f (x) = 6x 2
− 18x − 24

35. f (x) = x sin x on [−1, 1]


36. f (x) = x+1
x

37. f (x) = x+1


x

In Exercises 38-41, find the requested derivative.


38. f (x) = x sin x; find f
′′
(x) .
39. f (x) = x sin x; find f
(4)
(x) .
40. f (x) = csc x; find f
′′
(x) .
41. f (x) = (x 3
− 5x + 2)(x
2
+ x − 7); find f
(8)
(x) .
In Exercises 42-45, use the graph of f (x) to sketch f ′
.
(x)

42.

2.E.8 https://math.libretexts.org/@go/page/9969
43.

44.

45.

2.5: The Chain Rule


Terms and Concepts
1. T/F: The Chain Rule describes how to evaluate the derivative of a composition of functions.
2. T/F: The Generalized Power Rule states that d

dx
n
(g(x ) ) = n(g(x))
n−1
.
3. T/F: d

dx
(ln(x )) =
2 1

x2
.

4. T/F: d

dx
x
(3 ) ≈ 1.1 ⋅ 3
x
.

5. T/F: dx

dy
=
dx

dt

dt

dy

2.E.9 https://math.libretexts.org/@go/page/9969
6. T/F: Taking the derivative of f (x) = x 2
sin(5x) requires the use of both the Product and Chain Rules.

Problems
In Exercises 7-28, compute the derivatives of the given function.
1
7. f (x) = (4x 3
− x) 0

8. f (t) = (3t − 2) 5

9. g(θ) = (sin θ + cos θ) 3

10. h(t) − e
2
3 t +y−1

4
11. f (x) = (x + 1

x
)

12. f (x) = cos(3x)


13. g(x) = tan(5x)
14. h(t) = sin 4
(2t)

15. p(t) = cos 3


(t
2
+ 3t + 1)

16. f (x) = ln(cos x)


17. f (x) = ln(x 2
)

18. f (x) = 2 ln(x)


19. g(r) = 4 r

20. g(t) = 5 cos t

21. g(t) = 15 2

22. m(w) = 3

2
w

23. h(t) = 2 +3
t
3 +2

24. m(w) = 3

2
+1
w

2
x
3 +x
25. f (x) = x
2
2

26. f (x) = x 2
sin(5x)

27. g(t) = cos(t 2


+ 3t) sin(5t − 7)

28. g(t) = cos(


2
1 5t
)e
t

In Exercises 29-32, find the equation 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 7 through 10.
10
29. f (x) = (4x 3
− x) at x = 0

30. f (t) = (3t − 2) 5


at t = 1

3
31. g(θ) = (sin θ + cos θ) at θ = π/2

32. h(t) = e
2
3 t +t−1
at t = −1

33. Compute (ln(kx)) two ways:


dx
d

(a) Using the Chain rule, and


(b) by first using the logarithm rule ln(ab) = ln a + ln b , then taking the derivative.
34. Compute (ln(x )) two ways:
dx
d k

(a) Using the Chain Rule, and


(b) by first using the logarithm rule ln(a p
) = p ln a , then taking the derivative.

2.E.10 https://math.libretexts.org/@go/page/9969
Review
35. The “wind chill factor” is a measurement of how cold it “feels” during cold, windy weather. Let W (w) be the wind chill factor,
in degrees Fahrenheit, when it is 25 F outside with a wind of w mph.
(a) What are the units of W (w)? ′

(b) What would you expect the sign of W (10) to be? ′

36. Find the derivatives of the following functions.


(a) f (x) = x e cot x 2 x

(b) g(x) = 2 3 4 x x x

2.6: Implicit Differentiation


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. T/F: Implicit differentiation can be used to find the derivative of y = √−
x.

4. T/F: Implicit differentiation can be used to find the derivative of y = x 3/4


.

Problems
In Exercises 5-12, compute the derivative of the given function.
5. f (x) = √−
x+
1

√x

6. f (x) = √−
x +x
3 2/3

−−−−−
7. f (x) = √1 − t 2

8. g(t) = √t sin t
9. h(x) = x 1.5

10. f (x) = x π
+x
1.9

1.9

x+7
11. g(x) = √x

12. f (t) = √t (sec t + e 5 t


)

dy
In Exercises 13-25, find dx
using implicit differentiation.
13. x 4
+y
2
+y = 7

14. x 2/5
+y
2/5
=1

15. cos(x) + sin(y) = 1


16. f racxy = 10
y
17. x
= 10

18. x 2
e
2
+2
y
=5

19. x 2
tan y = 50

4
20. (3x 2
+ 2y )
3
=2

2
21. (y 2
+ 2y − x) = 200

2
x +y
22. x+y
2
= 17

sin(x)+y
23. cos(y)+x
=1

24. ln(x 2
+y ) = e
2

2.E.11 https://math.libretexts.org/@go/page/9969
25. ln(x 2
+ xy + y ) = 1
2

dy
26. Show that is the same for each of the following implicitly defined functions.
dx

(a) xy = 1
(b) x y = 1
2 2

(c) sin(xy) = 1
(d) ln(xy) = 1
In Exercises 27-31, find the equation of the tangent line to the graph of the implicitly defined function at the indicated
points. As a visual aid, each function is graphed.
27. x + y = 1
2/5 2/5

(a) At (1,0)
(b) At (0.1, 0.281) (which does not exactly lie on the curve, but is very close).

28. x + y = 1
4 4

(a) At (1,0).
−−− −
−−
(b) At (√0.6, √0.8).
(c) At (0,1).

29. (x + y − 4) = 108y
2 2 3 2

(a) At (0,4).
−−−
(b) At (2, −√108) 4

30. (x + y
2 2
+ x)
2
=x
2
+y
2

(a) At (0,1).

2.E.12 https://math.libretexts.org/@go/page/9969
3 √3
(b) At (− 3

4
,
4
) .

31. (x − 2) 2
+ (y − 3 )
2
=9

6+3 √3
(a) At ( 7

2
,
2
) .
4+3 √3
(b) At ( 2
,
3

2
) .

In Exercises 32-35, an implicitly defined function is given. Find d

dx
2
. Note: these are the same problems used in Exercises 13-
16.
32. x 4
+y
2
+y = 7

33. x 2/5
+y
2/5
=1

34. cos x + sin y = 1


35. x

y
= 10

dy
In Exercises 36-41, use logarithmic differentiation to find dx
, then find the equation of the tangent line at the indicated x-
value.
36. y = (1 + x) 1/x
, x =1

37. y = 2x
2
x
, x =1

38. y = x

x+1
, x =1

39. y = x sin(x)+2
, x =1

40. y = x+1

x+2
, x =1

(x+1)(x+2)
41. y = (x+3)(x+4)
, x =1

2.7: Derivatives of Inverse Functions


Terms and Concepts
1. T/F: Every function has an inverse.
2. In your own words explain what it means for a function to be "one on one."

2.E.13 https://math.libretexts.org/@go/page/9969
3. If (1,10) lies on the graph of y = f (x), what can be said about the graph of y = f −1
(x) ?
4. If (1,10) lies on the graph of y = f (x) and f ′
(1) = 5, what can be said about y = f −1
(x) ?

Problems
In Exercises 5-8, verify that the given functions are inverses.
5. f (x) = 2x + 6 and g(x) = 1

2
x −3

−−−−−
6. f (x) = x 2
+ 6x + 11, x ≥ 3 and g(x) = √x − 2 − 3, x ≥ 2

7. f (x) = 3

x−5
, x ≠5 and g(x) = 3+5x

x
, x ≠0

x+1
8. f (x) = x−1
, x ≠ 1 and g(x) = f (x)

In Exercises 9-14, an invertible function f (x) is given along with a point that lies on its graph. Using Theorem 22, evaluate

) (x) at the indicated value.
−1
(f

9. f (x) = 5x + 10
Point = (2,20)

Evaluate (f ) (20)
−1

10. f (x) = x 2
− 2x + 4, x ≥ 1

Point = (3, 7)

Evaluate (f −1
) (7)

11. f (x) = sin 2x, −π/4 ≤ x ≤ π/4



Point = (π/6, √3/2)
– ′
Evaluate (f ) (√3/2)
−1

12. f (x) = x 3
− 6x
2
+ 15x − 2

Point = (1, 8)

Evaluate (f −1
) (8)

13. f (x) = 1+x


1
2
, x ≥0

Point = (1, 1/2)



Evaluate (f ) (1/2)
−1

14. f (x) = 6e 3x

Point = (0, 6)

Evaluate (f ) (6)
−1

In Exercises 15-24, compute the derivative of the given function.


15. h(t) = sin −1
(2t)

16. f (t) = sec −1


(2t)

17. g(x) = tan −1


(2x)

18. f (x) = x sin −1


(x)

19. g(t) = sin t cos −1


t

20. f (t) = ln te t

−1

21. h(x) = sin

cos
−1
x


22. g(x) = tan −1
(√x )

23. f (x) = sec −1


(1/x)

24. f (x) = sin(sin −1


x)

In Exercises 25-27, compute the derivative of the given function in two ways:
(a) By simplifying first, then taking the derivative, and

2.E.14 https://math.libretexts.org/@go/page/9969
(b) by using the Chain Rule first then simplifying.
Very 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)

In Exercises 28-29, find the equation of the line tangent to the graph of f at the indicated value.
√2
28. f (x) = sin −1
x at x =
2

√3
29. f (x) = cos −1
(2x) at x =
4

Review
dy
30. Find dx
, where x 2 2
y −y x = 1 .
31. Find the equation of the line tangent to the graph of x 2
+y
2
+ xy = 7 at the point (1,2).
f (x+s)−f (x)
32. Let f (x) = x 3
+x . Evaluate lim s
.
s→0

2.E: Applications of Derivatives(Exercises) is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by LibreTexts.

2.E.15 https://math.libretexts.org/@go/page/9969
CHAPTER OVERVIEW
3: The Graphical Behavior of Functions
Our study of limits led to continuous functions, which is 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 information that continuous and differentiable functions provide.
3.1: Extreme Values
3.2: The Mean Value Theorem
3.3: Increasing and Decreasing Functions
3.4: Concavity and the Second Derivative
3.5: Curve Sketching
3.E: Applications of the Graphical Behavior of Functions(Exercises)

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 3: The Graphical Behavior of Functions is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by
Gregory Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available
upon request.

1
3.1: Extreme Values
Our study of limits led to continuous functions, which is 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 information that continuous and differentiable functions
provide.
Given any quantity described by a function, we are often interested in the largest and/or smallest values that quantity attains. For
instance, if a function describes the speed of an object, it seems reasonable to want to know the fastest/slowest the object traveled.
If a function describes the value of a stock, we might want to know how the highest/lowest values the stock attained over the past
year. We call such values extreme values.

Definition 3.1.1: Minima and Maxima

Let f be defined on an interval I containing c .


f (c) is the minimum (also, absolute minimum) of f on I if f (c) ≤ f (x) for all x in I .
f (c) is the maximum} (also, absolute maximum) of f on I if f (c) ≥ f (x) for all x in I .
The maximum and minimum values are the extreme values, or extrema, of f on I .

The extreme values of a function are "y '' values, values the function attains, not the input values.

Consider Figure 3.1.1. The function displayed in (a) has a maximum, but no minimum, as the interval over which the function is
defined is open. In (b), the function has a minimum, but no maximum; there is a discontinuity in the "natural'' place for the
maximum to occur. Finally, the function shown in (c) has both a maximum and a minimum; note that the function is continuous and
the interval on which it is defined is closed.

Figure 3.1.1: Graphs of functions with and without extreme values


It is possible for discontinuous functions defined on an open interval to have both a maximum and minimum value, but we have
just seen examples where they did not. On the other hand, continuous functions on a closed interval always have a maximum and
minimum value.

Theorem 3.1.1: The Extreme Value Theorem

Let f be a continuous function defined on a closed interval I . Then f has both a maximum and minimum value on I .

This theorem states that f has extreme values, but it does not offer any advice about how/where to find these values. The process
can seem to be fairly 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 values.

3.1.1 https://math.libretexts.org/@go/page/4164
Example 3.1.1: Approximating Extreme Values

Consider f (x) = 2x 3
− 9x
2
on I = [−1, 5] , as graphed in Figure 3.1.2. Approximate the extreme values of f .

Figure 3.1.2: A graph of f (x) = 2x 3


− 9x
2
as in Example 3.1.1
Solution
The graph is drawn in such a way to draw attention to certain points. It certainly seems that the smallest y value is −27, found
when 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 be sure the smallest value is not less than −27. Since the problem asks for an approximation, we
approximate the extreme values to be 25 and −27.

Notice how the minimum value came at "the bottom of a hill," and the maximum value came at an endpoint. Also note that while 0
is not an extreme value, it would be if we narrowed our interval to [−1, 4]. The idea that the point (0, 0) is the location of an
extreme value for some interval is important, leading us to a definition.

Local and Relative Extrema


The terms local minimum and local maximum are often used as synonyms for relative minimum and relative maximum. We
briefly practice using these definitions.

Definition 3.1.2: Relative Minimum and Relative Maximum

Let f be defined on an interval I containing c .


1. If there is an open interval containing c such that f (c) is the minimum value, 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 an open interval containing c such that f (c) is the maximum value, then f (c) is a relative maximum of f . We
also say that f has a relative maximum at (c, f (c)).
The relative maximum and minimum values comprise the relative extrema of f .

Example 3.1.2: Approximating Relative Extrema

Consider f (x) = (3x − 4x4 3


− 12 x
2
+ 5)/5 , as shown in Figure . Approximate the relative extrema of f . At each of
3.1.3

these points, evaluate f . ′

3.1.2 https://math.libretexts.org/@go/page/4164
Figure 3.1.3: A graph of f (x) = (3x 4
− 4x
3
− 12 x
2
+ 5)/5 as in Example 3.1.2.
Solution
We still do not have the tools to exactly find the relative extrema, but the graph does allow us to make reasonable
approximations. It seems f has relative minima at x = −1 and x = 2 , with values of f (−1) = 0 and f (2) = −5.4. It also
seems that f has a relative maximum at the point (0, 1).
We approximate the relative minima to be 0 and −5.4; we approximate the relative maximum to be 1.
It is straightforward to evaluate f ′
(x) =
1

5
3
(12 x − 12 x
2
− 24x) at x = 0, 1 and 2. In each case, f ′
(x) = 0 .

Example 3.1.3: Approximating Relative Extrema

Approximate the relative extrema of f (x) = (x − 1) 2/3


+2 , shown in Figure 3.1.4. At each of these points, evaluate f . ′

Figure 3.1.4: A graph of f (x) = (x − 1) 2/3


+2 as in Example 3.1.3.
Solution
The figure implies that f does not have any relative maxima, but has a relative minimum at (1, 2). In fact, the graph suggests
that not only is this point a relative minimum, y = f (1) = 2 the minimum value of the function.
We compute f ′
(x) =
2

3
−1/3
(x − 1 ) . When x = 1 , f is undefined.

What can we learn from the previous two examples? We were able to visually approximate relative extrema, and at each such point,
the derivative was either 0 or it was not defined. This observation holds for all functions, leading to a definition and a theorem.

Definition 3.1.3: Critical Numbers and Critical Points

Let f be defined at c . The value c is a critical number (or critical value) of f if f ′


(c) = 0 or f ′
(c) is not defined.
If c is a critical number of f , then the point (c, f (c)) is a critical point of f .

3.1.3 https://math.libretexts.org/@go/page/4164
Theorem 3.1.1: Relative Extrema and Critical Points

Let a function f have a relative extrema at the point (c, f (c)). Then c is a critical number of f .

Be careful to understand that this theorem states "All relative extrema occur at critical points." It does not say "All critical numbers
produce relative extrema." For instance, consider f (x) = x . Since f (x) = 3x , it is straightforward to determine that x = 0 is a
3 ′ 2

critical number of f . However, f has no relative extrema, as illustrated in Figure 3.1.5.

Figure 3.1.5: A graph of f (x) = x which has a critical value of x = 0 , but no relative extrema.
3

Theorem 3.1.1 states that a continuous function on a closed interval will have absolute extrema, that is, both an absolute maximum
and an absolute minimum. These extrema occur either at the endpoints or at critical values in the interval. We combine these
concepts to offer a strategy for finding extrema.

Key Idea 2: Finding Extrema on a Closed Interval


Let f be a continuous function defined on a closed interval [a, b]. To find the maximum and minimum values of f on [a, b]
1. Evaluate f at the endpoints a and b of the interval.
2. Find the critical numbers of f in [a, b].
3. Evaluate f at each critical number.
4. The absolute maximum of f is the largest of these values, and the absolute minimum of f is the least of these values.

We practice these ideas in the next examples.

Example 3.1.4: Finding Extreme Values

Find the extreme values of f (x) = 2x 3 2


+ 3x − 12x on [0, 3], graphed in Figure 3.1.6.
We follow the steps outlined in the Key Idea 2. We first evaluate f at the endpoints:

f (0) = 0 and f (3) = 45. (3.1.1)

Figure 3.1.6: A graph of f (x) = 2x 3


+ 3x
2
− 12x on [0, 3] as in Example 3.1.4.

3.1.4 https://math.libretexts.org/@go/page/4164
Next, we find the critical values of f on [0, 3]. f (x) = 6x + 6x − 12 = 6(x + 2)(x − 1) ; therefore the critical values of f
′ 2

are x = −2 and x = 1 . Since x = −2 does not lie in the interval [0, 3], we ignore it. Evaluating f at the only critical number
in our interval gives: f (1) = −7 .
Table 3.1.1 gives f evaluated at the "important" x values in [0, 3]. We can easily see the maximum and minimum values of f :
the maximum value is 45 and the minimum value is −7.
Table 3.1.1 : Finding the extreme values of f in Example 3.1.4
x f(x)

0 0

1 -7

3 45

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 3.1.6 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.

Example 3.1.5: Finding Extreme Values

Find the maximum and minimum values of f on [−4, 2], where


2
(x − 1) x ≤0
f (x) = { . (3.1.2)
x +1 x >0

Solution
Here f is piecewise--defined, but we can still apply the Key Idea 2. Evaluating f at the endpoints gives:
f (−4) = 25 and f (2) = 3. (3.1.3)

We now find the critical numbers of f . We have to define f in a piecewise manner; it is


2(x − 1) x <0

f (x) = { . (3.1.4)
1 x >0

Note that while f is defined for all of [−4, 2], f is not, as the derivative of f does not exist when x = 0 . (From the left, the

derivative approaches −2; from the right the derivative is 1.) Thus one critical number of f is x = 0 .
We now set f (x) = 0 . When x > 0 , f (x) is never 0. When x < 0 , f (x) is also never 0. (We may be tempted to say that
′ ′ ′

f (x) = 0 when x = 1 . However, this is nonsensical, for we only consider f (x) = 2(x − 1) when x < 0 , so we will ignore a
′ ′

solution that says x = 1 .)


So we have three important x values to consider: x = −4, 2 and 0. Evaluating f at each gives, respectively, 25, 3 and 1,
shown in Table 3.1.2. Thus the absolute minimum of f is 1; the absolute maximum of f is 25. Our answer is confirmed by the
graph of f in Figure 3.1.7.
Table 3.1.2 : Finding the extreme values of f in Example 3.1.5
x f (x)

-4 25

0 1

2 3

3.1.5 https://math.libretexts.org/@go/page/4164
Figure 3.1.7: A graph of f (x) on [−4, 2] as in Example 3.1.5.

Example 3.1.6: Finding Extreme Values

Find the extrema of f (x) = cos(x 2


) on [−2, 2].
Solution
We again use Key Idea 3. Evaluating f at the endpoints of the interval gives: f (−2) = f (2) = cos(4) ≈ −0.6536. We now
find the critical values of f .
Applying the Chain Rule, we find f ′
(x) = −2x sin(x )
2
. Set f ′
(x) = 0 and solve for x to find the critical values of f .
We have ′
f (x) = 0 when and when sin(x ) = 0 . In general, sin t = 0 when t = … − 2π, −π, 0, π, … Thus
x =0
2

2
sin(x ) = 0 when 2
x (x is always positive so we ignore −π, etc.) So sin(x ) = 0
= 0, π, 2π, …
2
when 2

− −− − −
x = 0, ±√π , ±√2π, …. The only values to fall in the given interval of [−2, 2] are −√π and √π , approximately ±1.77.

We again construct a table of important values in Table 3.1.3. In this example we have 5 values to consider: x = 0, ±2, ±√−
π.

Table 3.1.3 : Finding the extrema of f (x) = cos(x ) in Example 3.1.5


2

x f(x)

-2 -0.65

−√π -1

0 1

√π -1

2 -0.65

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 3.1.8

confirms our results.

Figure 3.1.8: A graph of f (x) = cos(x 2


) on [−2, 2] as in Example 3.1.5.

We consider one more example.

3.1.6 https://math.libretexts.org/@go/page/4164
Example 3.1.7: Finding Extreme Values
−−−−−
Find the extreme values of f (x) = √1 − x . 2

Solution
A closed interval is not given, so we find the extreme values of f on its domain. f is defined whenever 1 −x
2
≥0 ; thus the
domain of f is [−1, 1]. Evaluating f at either endpoint returns 0.
Using the Chain Rule, we find f ′
(x) =
−x
. The critical points of f are found when f ′
(x) = 0 or when f is undefined. It

√1−x2

is straightforward to find that f (x) = 0 when x = 0 , and



f

is undefined when x = ±1 , the endpoints of the interval. The
table of important values is given in Table 3.1.4.
Table 3.1.4 : Finding the extrema of the half--circle in Example 3.1.7
x f(x)

-1 0

0 1

1 0

The maximum value is 1, and the minimum value is 0.

−−−−−
Figure 3.1.9: A graph of f (x) = √1 − x on [−1, 1] as in Example 3.1.7
2

Note: We implicitly found the derivative of x


2
+y
2
=1 , the unit circle, in the section on Implicit Differentiation as
dy −−−−− −x

dx
= −x/y . In Example 3.1.7, half of the unit circle is given as y = f (x) = √1 − x . We found f 2 ′
(x) = . Recognize
√1−x2

dy
that the denominator of this fraction is y ; that is, we again found f ′
(x) =
dx
= −x/y.

We have seen that continuous functions on closed intervals always have a maximum and minimum value, and we have also
developed a technique to find these values. In the next section, we further our study of the information we can glean from "nice"
functions with the Mean Value Theorem. On a closed interval, we can find the average rate of change of a function (as we did at
the beginning of Chapter 2). We will see that differentiable functions always have a point at which their instantaneous rate of
change is same as the average rate of change. This is surprisingly useful, as we'll see.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/
Integrated by Justin Marshall.

This page titled 3.1: Extreme Values is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et
al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

3.1.7 https://math.libretexts.org/@go/page/4164
3.2: The Mean Value Theorem
We motivate this section with the following question: Suppose you leave your house and drive to your friend's house in a city 100
miles away, completing the trip in two hours. At any point during the trip do you necessarily have to be going 50 miles per hour?
In answering this question, it is clear that the average speed for the entire trip is 50 mph (i.e. 100 miles in 2 hours), but the question
is whether or not your instantaneous speed is ever exactly 50 mph. More simply, does your speedometer ever read exactly 50
mph?. The answer, under some very reasonable assumptions, is "yes."'
Let's now see why this situation is in a calculus text by translating it into mathematical symbols.
First assume that the function y = f (t) gives the distance (in miles) traveled from your home at time t (in hours) where 0 ≤ t ≤ 2 .
In particular, this gives f (0) = 0 and f (2) = 100. The slope of the secant line connecting the starting and ending points (0, f (0))
and (2, f (2)) is therefore
$$
\frac{\Delta f}{\Delta t} = \frac{f(2)-f(0)}{2-0} = \frac{100-0}{2} = 50 \, \text{mph}.
\]
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'(c) = \frac{f(2)-f(0)}{2-0} = 50
\]
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 (b) − f (a)

f (c) = (3.2.1)
b −a

have to hold for some a < c < b ?


Let's look at two functions in an example.

Example 3.2.1: Comparing average and instantaneous rates of change

Consider functions
$$f_1(x)=\frac{1}{x^2}\quad \text{and} \quad f_2(x) = |x|\]
with a = −1 and b = 1 as shown in Figure 3.2.1 (a) and (b), respectively. 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'(c) = \frac{f(1)-f(-1)}{1-(-1)} = 0.\]

3.2.1 https://math.libretexts.org/@go/page/4165
Figure 3.2.1 : A graph of f 1 (x) = 1/ x
2
and f
2 (x) = |x| in Example 3.2.1.

So what went "wrong"'? It may not be surprising to find that the discontinuity of f and the corner of f play a role. If our
1 2

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.

Theorem 3.2.1: The Mean Value Theorem of Differentiation

Let y = f (x) be continuous function on the closed interval [a, b] and differentiable on the open interval (a, b) . There exists a
value c , a < c < b , such that
f (b) − f (a)

f (c) = . (3.2.2)
b −a

That is, there is a value c in (a, b) where the instantaneous rate of change of f at c is equal to the average rate of change of f on
[a, b].

Note that the reasons that the functions in Example 3.2.1 fail are indeed that f has a discontinuity on the interval [−1, 1] and f is
1 2

not differentiable at the origin.


We will give a proof of the Mean Value Theorem below. To do so, we use a fact, called Rolle's Theorem, stated here.

Theorem 3.2.2: Rolle's Theorem

Let f be continuous on [a, b] and differentiable on (a, b), where f (a) = f (b) . There is some c in (a, b) such that f ′
(c) = 0.

Consider Figure 3.2.2 where the graph of a function f is given, where f (a) = f (b) . It should make intuitive sense that if f is
differentiable (and hence, continuous) that there would be a value c in (a, b) where f (c) = 0 ; that is, there would be a relative

maximum or minimum of f in (a, b). Rolle's Theorem guarantees at least one; there may be more.

FIgure 3.2.2 : A graph of f (x) = x 3 2


− 5x + 3x + 5 , where f (a) = f (b) . Note the existence of c , where a < c < b , where
f (c) = 0 .

3.2.2 https://math.libretexts.org/@go/page/4165
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.

Proof of Rolle's Theorem

Let f be differentiable on (a, b) where f (a) = f (b) . We consider two cases.


Case 1: Consider the case when f is constant on [a, b]; that is, f (x) = f (a) = f (b) for all x in [a, b]. Then f ′
(x) = 0 for all
x in [a, b], showing there is at least one value c in (a, b) where f (c) = 0 .

Case 2: Now assume that f is not constant on [a, b]. The Extreme Value Theorem guarantees that f has a maximal and
minimal value on [a, b], found either at the endpoints or at a critical value in (a, b). Since f (a) = f (b) and f is not constant, it
is clear that the maximum and minimum cannot both be found at the endpoints. Assume, without loss of generality, that the
maximum of f is not found at the endpoints. Therefore there is a c in (a, b) such that f (c) is the maximum value of f . By
Theorem 3.1.2, c must be a critical number of f ; since f is differentiable, we have that f (c) = 0 , completing the proof of the

theorem.

We can now prove the Mean Value Theorem.

Proof of the Mean Value Theorem


Define the function
$$g(x) = f(x) - \frac{f(b)-f(a)}{b-a}x.\]
We know g is differentiable on (a, b) and continuous on [a, b] since f is. We can show g(a) = g(b) (it is actually easier to
show g(b) − g(a) = 0 , which suffices). We can then apply Rolle's theorem to guarantee the existence of c ∈ (a, b) such that
g (c) = 0 . But note that

$$0= g'(c) = f'(c) - \frac{f(b)-f(a)}{b-a} \ ;\]


hence
$$f'(c) = \frac{f(b)-f(a)}{b-a},\]
which is what we sought to prove.

Going back to the very beginning of the section, we see that the only assumption we would need about our distance function f (t) is
that it be continuous and differentiable for t from 0 to 2 hours (both reasonable assumptions). By the Mean Value Theorem, we are
guaranteed a time during the trip where our instantaneous speed is 50 mph. This fact is used in practice. Some law enforcement
agencies monitor traffic speeds while in aircraft. They do not measure speed with radar, but rather by timing individual cars as they
pass over lines painted on the highway whose distances apart are known. The officer is able to measure the average speed of a car
between the painted lines; if that average speed is greater than the posted speed limit, the officer is assured that the driver exceeded
the speed limit at some time.
Note that the Mean Value Theorem is an existence theorem. It states that a special value c exists, but it does not give any indication
about how to find it. It turns out that when we need the Mean Value Theorem, existence is all we need

Example 3.2.2: Using the Mean Value Theorem

Consider f (x) = x 3
+ 5x + 5 on [−3, 3]. Find c in [−3, 3] that satisfies the Mean Value Theorem.
Solution
The average rate of change of f on [−3, 3] is:
f (3) − f (−3) 84
= = 14. (3.2.3)
3 − (−3) 6

3.2.3 https://math.libretexts.org/@go/page/4165
We want to find c such that f ′
(c) = 14 . We find f ′
(x) = 3 x
2
+5 . We set this equal to 14 and solve for x.

f (x) = 14

2
3x + 5 = 14
2
x =3

x = ±√3 ≈ ±1.732

We have found 2 values c in [−3, 3] where the instantaneous rate of change is equal to the average rate of change; the Mean
Value Theorem guaranteed at least one. In Figure 3.2.3 f is graphed with a dashed line representing the average rate of

change; the lines tangent to f at x = ±√3 are also given. Note how these lines are parallel (i.e., have the same slope) as the
dashed line.

Figure 3.2.3 : Demonstrating the Mean Value Theorem in Example 3.2.2.

While the Mean Value Theorem has practical use (for instance, the speed monitoring 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.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/
Integrated by Justin Marshall.

This page titled 3.2: The Mean Value Theorem is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory
Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

3.2.4 https://math.libretexts.org/@go/page/4165
3.3: Increasing and Decreasing Functions
Our study of "nice" functions f in this chapter has so far focused on individual 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 interval.
′ ′ ′

In this section we begin to study how functions behave between special points; we begin studying in more detail the shape of their graphs.
We start with an intuitive concept. Given the graph in Figure 3.3.1, where would you say the function is increasing? Decreasing?

Figure 3.3.1 : A graph of a function f used to illustrate the concepts of increasing and decreasing.
Even though we have not defined these terms mathematically, one likely answered that f is increasing when x >1 and decreasing when
x < 1 . We formally define these terms here.

Definition: Increasing and Decreasing Functions


Let f be a function defined on an interval I .\index{increasing function}\index{decreasing function}\index{increasing
function!strictly}\index{decreasing function!strictly}
1. f is increasing on I if for every a < b in I , f (a) ≤ f (b) .
2. f is decreasing on I if for every a < b in I , f (a) ≥ f (b) .
A function is strictly increasing when a < b in I implies f (a) < f (b) , with a similar definition holding for strictly decreasing.

Informally, a function is increasing if as x gets larger (i.e., looking left to right) f (x) gets larger.
Our interest lies in finding intervals in the domain of f on which f is either increasing or decreasing. Such information should seem useful.
For instance, if f describes the speed of an object, we might want to know when the speed was increasing or decreasing (i.e., when the
object was accelerating vs. decelerating). If f describes the population of a city, we should be interested in when the population is growing
or declining.
To find such intervals, we again consider secant lines. Let f be an increasing, differentiable function on an open interval I , such as the one
shown in Figure 3.3.2, and let a < b be given in I . The secant line on the graph of f from x = a to x = b is drawn; it has a slope of
(f (b) − f (a))/(b − a) . But note:

f (b) − f (a) numerator > 0


⇒ ⇒ slope of the secent line > 0 ⇒ Average rate of chjange of f on [a, b] is > 0. (3.3.1)
b −a denominator > 0

Figure 3.3.2 : Examining the secant line of an increasing function.


We have shown mathematically what may have already been obvious: when f is increasing, its secant lines will have a positive slope. Now
recall the Mean Value Theorem guarantees that there is a number c , where a < c < b , such that

3.3.1 https://math.libretexts.org/@go/page/4166
f (b) − f (a)

f (c) = > 0. (3.3.2)
b −a

By considering all such secant lines in I , we strongly imply that f ′


(x) ≥ 0 on I . A similar statement can be made for decreasing functions.
Our above logic can be summarized as "If f is increasing, then f is probably positive." Theorem 3.3.1 below turns this around by stating

"If f is postive, then f is increasing." This leads us to a method for finding when functions are increasing and decreasing.

THeorem 3.3.1: Test For Increasing/Decreasing Functions

Let f be a continuous function on [a, b] and differentiable on (a, 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].

Note: Theorem 3.3.1 also holds if f ′


(c) = 0 for a finite number of values of c in I .
Let a and b be in I where f (a) > 0 and f (b) < 0 . It follows from the Intermediate Value Theorem that there must be some value c
′ ′

between a and b where f (c) = 0 . This leads us to the following method for finding intervals on which a function is increasing or

decreasing.

Key Idea 3: Finding Intervals on Which f is Increasing or Decreasing

Let f be a differentiable function on an interval I. To find intervals on which f is increasing and decreasing:
1. Find the critical values of f . That is, find all c in I where f (c) = 0 or f is not defined.
′ ′

2. Use the critical values to divide I into subintervals.


3. Pick any point p in each subinterval, and find the sign of f (p). ′

a. If f ′
(p) > 0 , then f is increasing on that subinterval.
b. If f ′
(p) < 0 , then f is decreasing on that subinterval.

We demonstrate using this process in the following example.

Example 3.3.1: Finding intervals of increasing/decreasing

Let f (x) = x 3
+x
2
−x +1 . Find intervals on which f is increasing or decreasing.
Solution
Using the Key Idea 3, we first find the critical values of f . We have ′
f (x) = 3 x
2
+ 2x − 1 = (3x − 1)(x + 1) , so ′
f (x) = 0 when
x = −1 and when x = 1/3. f is never undefined.

Since an interval was not specified for us to consider, we consider the entire domain of f which is (−∞, ∞). We thus break the whole
real line into three subintervals based on the two critical values we just found: (−∞, −1), (−1, 1/3) and (1/3, ∞). This is shown in
Figure 3.3.3.

Figure 3.3.3 : Number line for f in Example 3.3.1


We now pick a value p in each subinterval and find the sign of f ′
. All we care about is the sign, so we do not actually have to fully
(p)

compute f (p); pick "nice" values that make this simple.


Subinterval 1, (−∞, −1): We (arbitrarily) pick p = −2 . We can compute ′


f (−2) directly: f ′
(−2) = 3(−2 )
2
+ 2(−2) − 1 = 7 > 0 .
We conclude that f is increasing on (−∞, −1).
Note we can arrive at the same conclusion without computation. For instance, we could choose p = −100. The first term in f (−100), ′

i.e., 3(−100) is clearly positive and very large. The other terms are small in comparison, so we know f (−100) > 0. All we need is
2 ′

the sign.
Subinterval 2, (−1, 1/3): We pick p = 0 since that value seems easy to deal with. ′
f (0) = −1 < 0 . We conclude f is decreasing on
(−1, 1/3).

3.3.2 https://math.libretexts.org/@go/page/4166
Subinterval 3, (1/3, ∞): Pick an arbitrarily large value for p > 1/3 and note that ′ 2
f (p) = 3 p + 2p − 1 > 0 . We conclude that f is
increasing on (1/3, ∞).
We can verify our calculations by considering Figure 3.3.4, where f is graphed. The graph also presents f ; note how f ′ ′
>0 when f is
increasing and f < 0 when f is decreasing.

Figure 3.3.4 : A graph of f (x) in Example 3.3.1, showing where f is increasing and decreasing.

One is justified in wondering why so much work is done when the graph seems to make the intervals very clear. We give three reasons why
the above work is worthwhile.
First, the points at which f switches from increasing to decreasing are not precisely known given a graph. The graph shows us something
significant happens 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 the significant points are x = −1 and x = 1/3, the graph
shows the increasing/decreasing traits just fine. That is true. However, the technique prescribed here helps reinforce the relationship between
increasing/decreasing and the sign of f . Once mastery of this concept (and several others) is obtained, one finds that either (a) just the

critical points are computed and the graph shows all else that is desired, or (b) a graph is never produced, because determining
increasing/decreasing using f is straightforward and the graph is unnecessary.

So our second reason why the above work is worthwhile is this: once mastery of a subject is gained, one has options for finding needed
information. We are working to develop mastery.
Finally, our third reason: many problems we face "in the real world" are very complex. Solutions are tractable only through the use of
computers to do many calculations for us. Computers do not solve problems "on their own," however; they need to be taught (i.e.,
programmed) to do the right things. It would be beneficial to give a function to a computer and have it return maximum and minimum
values, intervals on which the function is increasing and decreasing, the locations of relative maxima, etc. The work that we are doing here
is easily programmable. It is hard to teach a computer to "look at the graph and see if it is going up or down." It is easy to teach a computer
to "determine if a number is greater than or less than 0."
In Section 3.1 we learned the definition of relative maxima and minima and found that they occur at critical points. We are now learning that
functions can switch from increasing to decreasing (and vice--versa) at critical points. This new understanding of increasing and decreasing
creates a great method of determining whether a critical point corresponds to a maximum, minimum, or neither. Imagine a function
increasing until a critical point at x = c , after which it decreases. A quick sketch helps confirm that f (c) must be a relative maximum. A
similar statement can be made for relative minimums. We formalize this concept in a theorem.

THeorem 3.3.2: First Derivative Test

Let f be differentiable on I and let c be a critical number in I .


1. If the sign of f switches from positive to negative at c , then f (c) is a relative maximum of f .

2. If the sign of f switches from negative to positive at c , then f (c) is a relative minimum of f .

3. If the sign of f does not change at c , then f (c) is not a relative extrema of f .

Example 3.3.2: Using the First Derivative Test

Find the intervals on which f is increasing and decreasing, and use the First Derivative Test to determine the relative extrema of f ,
where
2
x +3
f (x) = . (3.3.3)
x −1

Solution

3.3.3 https://math.libretexts.org/@go/page/4166
We start by noting the domain of f : (−∞, 1) ∪ (1, ∞) . Key Idea 3 describes how to find intervals where f is increasing and decreasing
when the domain of f is an interval. Since the domain of f in this example is the union of two intervals, we apply the techniques of Key
Idea 3 to both intervals of the domain of f .
Since f is not defined at x = 1 , the increasing/decreasing nature of f could switch at this value. We do not formally consider x = 1 to
be a critical value of f , but we will include it in our list of critical values that we find next.
Using the Quotient Rule, we find
2

x − 2x − 3
f (x) = . (3.3.4)
(x − 1)2

We need to find the critical values of f ; we want to know when 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.
′ ′

Note: Strictly speaking, x = 1 is not a critical value of f as f is not defined at x = 1 . We therefore actually apply Key Idea 3 to the
intervals (−∞, 1) and (1, ∞). We make note of x = 1 on the number line as we recognize that the behavior of f can change there, as it
is not defined there.

f (x) = 0 when the numerator of f ′
(x) is 0. That occurs when x 2
− 2x − 3 = (x − 3)(x + 1) = 0 ; i.e., when x = −1, 3.
We have found that f has two critical numbers, x = −1, 3, and at x = 1 something important might also happen. These three numbers
divide the real number line into 4 subintervals:
(−∞, −1), (−1, 1), (1, 3) and (3, ∞). (3.3.5)

Pick a number p from each subinterval and test the sign of f at p to determine whether f is increasing or decreasing on that interval.

Again, we do well to avoid complicated computations; notice that the denominator of f is always positive so we can ignore it during

our work.
Interval 1, (−∞, −1): Choosing a very small number (i.e., a negative number with a large magnitude) p returns 2
p − 2p − 3 in the
numerator of f ; that will be positive. Hence f is increasing on (−∞, −1).

Interval 2, (−1, 1): Choosing 0 seems simple: f ′


(0) = −3 < 0 . We conclude f is decreasing on (−1, 1).
Interval 3, (1, 3): Choosing 2 seems simple: f ′
(2) = −3 < 0 . Again, f is decreasing.
Interval 4, (3, ∞): Choosing an very large number p from this subinterval will give a positive numerator and (of course) a positive
denominator. So f is increasing on (3, ∞).

Figure 3.3.5 : Number line for f in Example 3.3.2


In summary, f is increasing on the set (−∞, −1) ∪ (3, ∞) and is decreasing on the set (−1, 1) ∪ (1, 3). Since at x = −1 , the sign of
f'\ switched from positive to negative, Theorem 3.3.2 states that f (−1) is a relative maximum of f . At x = 3 , the sign of f'\ switched
from negative to positive, meaning f (3) is a relative minimum. At x = 1 , f is not defined, so there is no relative extrema at x = 1 .

Figure 3.3.6 : A graph of f (x) in Example 3.3.2, showing where f is increasing and decreasing.
This is summarized in the number line shown in Figure 3.3.3. Also, Figure 3.3.4 shows a graph of f , confirming our calculations. This
figure also shows f , again demonstrating that f is increasing when f > 0 and decreasing when f < 0 .
′ ′ ′

3.3.4 https://math.libretexts.org/@go/page/4166
One is often tempted to think that functions always alternate "increasing, decreasing, 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.

Example 3.3.3: Using the First Derivative Test

Find the intervals on which f (x) = x 8/3


− 4x
2/3
is increasing and decreasing and identify the relative extrema.
Solution
We again start with taking derivatives. Since we know we want to solve f ′
(x) = 0 , we will do some algebra after taking derivatives.
8 2

f (x) = x 3
− 4x 3
(3.3.6)

8 5 8 −
1

f (x) = x 3 − x 3 (3.3.7)
3 3
8 1 6

= x 3 (x 3 − 1) (3.3.8)
3
8 −
1
2
= x 3
(x − 1) (3.3.9)
3
8 −
1

= x 3
(x − 1)(x + 1). (3.3.10)
3

This derivation of f shows that f (x) = 0 when x = ±1 and f is not defined when x = 0 . Thus we have 3 critical values, breaking
′ ′ ′

the number line into 4 subintervals as shown in Figure 3.3.5.


Interval 1, (∞, −1): We choose p = −2 ; we can easily verify that f ′
(−2) < 0 . So f is decreasing on (−∞, −1).
Interval 2, (−1, 0): Choose p = −1/2. Once more we practice finding the sign of f ′
(p) without computing an actual value. We have
(p − 1)(p + 1) ; find the sign of each of the three terms.
′ −1/3
$ f (p) = (8/3)p

8 −
1

f (p) = ⋅p 3 ⋅ (p − 1) (p + 1) . (3.3.11)
3   
<0 <0 >0

We have a "negative × negative × positive" giving a positive number; f is increasing on (−1, 0).
Interval 3, (0, 1): We do a similar sign analysis as before, using p in (0, 1).


8 −
1

f (p) = ⋅ p 3 ⋅ (p − 1) (p + 1) . (3.3.12)
3   
>0 <0 >0

We have 2 positive factors and one negative factor; f ′


(p) < 0 and so f is decreasing on (0, 1).

Interval 4, (1, ∞): Similar work to that done for the other three intervals shows that ′
f (x) > 0 on (1, ∞) , so f is increasing on this
interval.

Figure 3.3.7 : Number line for f in Example 3.3.3


We conclude by stating that f is increasing on (−1, 0) ∪ (1, ∞) and decreasing on (−∞, −1) ∪ (0, 1) . The sign of f changes from ′

negative to positive around x = −1 and x = 1 , meaning by Theorem 3.3.2 that f (−1) and f (1) are relative minima of f . As the sign
of f changes from positive to negative at x = 0 , we have a relative maximum at f (0). Figure 3.3.8 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 .
′ ′ ′

3.3.5 https://math.libretexts.org/@go/page/4166
Figure 3.3.8 : A graph of f (x) in Example 3.3.3, showing where f is increasing and decreasing.

We have seen how the first derivative of a function helps determine when the 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.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of VMI and
Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution - Noncommercial (BY-
NC) License. http://www.apexcalculus.com/
Integrated by Justin Marshall.

This page titled 3.3: Increasing and Decreasing Functions is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory
Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

3.3.6 https://math.libretexts.org/@go/page/4166
3.4: Concavity and the Second Derivative
Our study of "nice" functions continues. The previous section showed how the first derivative of a function, f , can relay important

information about f . We now apply the same technique to f itself, and learn what this tells us about f . The key to studying f is to
′ ′

consider its derivative, namely f , which is the second derivative of f . When f > 0 , f is increasing. When f < 0 , f is
′′ ′′ ′ ′′ ′

decreasing. f has relative maxima and minima where f = 0 or is undefined. This section explores how knowing information
′ ′′

about f gives information about f .


′′

Concavity
We begin with a definition, then explore its meaning.

Definition Concave Up and Concave Down


Let f be differentiable on an interval I . The graph of f is concave up on I if f is increasing. The graph of f is concave down

on I if f is decreasing. If f is constant then the graph of f is said to have no concavity.


′ ′

Note: We often state that "f is concave up" instead of "the graph of f is concave up" for simplicity.
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 will be increasing. Consider Figure 3.4.1, where a concave up graph is shown along with some
tangent lines. Notice how the tangent line on the left is steep, downward, corresponding to a small value of f . On the right, the

tangent line is steep, upward, corresponding to a large value of f . ′

Figure 3.4.1 : A function f with a concave up graph. Notice how the slopes of the tangent lines, when looking from left to right,
are increasing.
If a function is decreasing and concave up, then its rate of decrease is slowing; it is "leveling off." If the function is increasing and
concave up, then the rate of increase is increasing. The function is increasing at a faster and faster rate.
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 lines will be decreasing. Consider Figure 3.4.2, where a concave down graph is shown along with
some tangent lines. Notice how the tangent line on the left is steep, upward, corresponding to a large value of f . On the right, the

tangent line is steep, downward, corresponding to a small value of f . ′

3.4.1 https://math.libretexts.org/@go/page/4167
Figure 3.4.2 : A function f with a concave down graph. Notice how the slopes of the tangent lines, when looking from left to right,
are decreasing.
If a function is increasing and concave down, then its rate of increase is slowing; 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 rate.
Note: A mnemonic for remembering what concave up/down means is: "Concave up is like a cup; concave down is like a frown." It
is admittedly terrible, but it works.
Our definition of concave up and concave down is given in terms of when the first derivative is increasing or decreasing. We can
apply the results of the previous section and to find intervals on which a graph is concave up or down. That is, we recognize that f ′

is increasing when f > 0 , etc.


′′

THeorem 3.4.1: 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.

Figure 3.4.3 : Demonstrating the 4 ways that concavity interacts with increasing/decreasing, along with the relationships with the
first and second derivatives.
Note: Geometrically speaking, a function is concave up if its graph lies above its tangent lines. A function is concave down if its
graph lies below its tangent lines.
If knowing where a graph is concave up/down is important, it makes sense that the places where the graph changes from one to the
other is also important. This leads us to a definition.

3.4.2 https://math.libretexts.org/@go/page/4167
Definition: Point of Inflection
A point of inflection is a point on the graph of f at which the concavity of f changes.

Figure 3.4.4 shows a graph of a function with inflection points labeled.

Figure 3.4.4 : A graph of a function with its inflection points marked. The intervals where concave up/down are also indicated.
If the concavity of f changes at a point (c, f (c)), then f is changing from increasing to decreasing (or, decreasing to increasing) at

x = c . That means that the sign of f is changing from positive to negative (or, negative to positive) at x = c . This leads to the
′′

following theorem.

THeorem 3.4.2: Points of Inflection

If (c, f (c)) is a point of inflection on the graph of f , then either f ′′


=0 or f is not defined at c .
′′

We have identified the concepts of concavity and points of inflection. It is 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.

Example 3.4.1: Finding intervals of concave up/down, inflection points

Let f (x) = x 3
− 3x + 1 . Find the inflection points of f and the intervals on which it is concave up/down.
Solution
We start by finding f (x) = 3x − 3 and f (x) = 6x . To find the inflection points, we use Theorem 3.4.2 and find where
′ 2 ′′

f (x) = 0 or where f is undefined. We find f is always defined, and is 0 only when x = 0 . So the point (0, 1) is the only
′′ ′′ ′′

possible point of inflection.


This possible inflection point divides the real line into two intervals, (−∞, 0) and (0, ∞). We use a process similar to the one
used in the previous section to determine increasing/decreasing. Pick any c < 0 ; f (c) < 0 so f is concave down on (−∞, 0).
′′

Pick any c > 0 ; f (c) > 0 so f is concave up on (0, ∞). Since the concavity changes at x = 0 , the point (0, 1) is an
′′

inflection point.

Figure 3.4.5 : A number line determining the concavity of f in Example 3.4.1.


The number line in Figure 3.4.5 illustrates the process of determining concavity; Figure 3.4.6 shows a graph of f and f
′′
,
confirming our results. Notice how f is concave down precisely when f (x) < 0 and concave up when f (x) > 0 .
′′ ′′

3.4.3 https://math.libretexts.org/@go/page/4167
Figure 3.4.6 : A graph of f (x) used in Example3.4.1

Example 3.4.2: Finding intervals of concave up/down, inflection points

Let f (x) = x/(x 2


− 1) . Find the inflection points of f and the intervals on which it is concave up/down.
Solution
We need to find f and f . Using the Quotient Rule and simplifying, we find
′ ′′

2 2
−(1 + x ) 2x(x + 3)
′ ′′
f (x) = and f (x) = . (3.4.1)
2 2 2 3
(x − 1) (x − 1)

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(x + 3) = 0 ; we find x = 0 . We find that f is not defined when x = ±1 , for then the denominator of
2 ′′

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.
The important x-values at which concavity might switch are x = −1 , x = 0 and x = 1 , which split the number line into four
intervals as shown in Figure 3.4.7. We determine the concavity on each. Keep in mind that all we are concerned with is the
sign of f on the interval.
′′

Interval 1, (−∞, −1): Select a number c in this interval with a large magnitude (for instance, c = −100 ). The denominator of
f (x) will be positive. In the numerator, the (c + 3) will be positive and the 2c term will be negative. Thus the numerator is
′′ 2

negative and f (c) is negative. We conclude f is concave down on (−∞, −1).


′′

Interval 2, (−1, 0): For any number c in this interval, the term 2c in the numerator will be negative, the term (c + 3) in the 2

numerator will be positive, and the term (c − 1) in the denominator will be negative. Thus f (c) > 0 and f is concave up
2 3 ′′

on this interval.
Interval 3, (0, 1): Any number c in this interval will be positive and "small." Thus the numerator is positive while the
denominator is negative. Thus f (c) < 0 and f is concave down on this interval.
′′

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, ∞).

Figure 3.4.7 : Number line for f in Example 3.4.2


We conclude that f is concave up on (−1, 0) ∪ (1, ∞) and concave down on (−∞, −1) ∪ (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 of f in Figure 3.4.8. Notice how f is

3.4.4 https://math.libretexts.org/@go/page/4167
concave up whenever f is positive, and concave down when f is negative.
′′ ′′

Figure 3.4.8 : A graph of f (x) and f ′′


(x) in Example 3.4.2

Recall that relative maxima and minima of f are found at critical points of 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 undefined; note that these
′ ′′ ′′

are the inflection points of f .


What does a "relative maximum of f " mean? The derivative measures the rate of change of f ; maximizing f means finding the
′ ′

where f is increasing the most -- where f has the steepest tangent line. A similar statement can be made for minimizing f ; it ′

corresponds to where f has the steepest negatively--sloped tangent line.


We utilize this concept in the next example.

Example 3.4.3: Understanding inflection points

The sales of a certain product over a three-year span are modeled by S(t) = t − 8t + 20 , where t is the time in years,
4 2

shown in Figure 3.4.9. Over the first two years, sales are decreasing. Find the point at which sales are decreasing at their
greatest rate.

Figure 3.4.9 : A graph of S(t) in Example 3.4.3, modeling the sale of a product over time.
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
′′
is 0. We find S (t) = 4t − 16t and S (t) = 12t − 16 . Setting S (t) = 0 and solving, we get t = √4/3 ≈ 1.16 (we
′ 3 ′′ 2 ′′

ignore the negative value of t since it does not lie in the domain of our function 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, it will still take some time before sales start to increase, but at least sales are not
decreasing quite as quickly as they had been.

3.4.5 https://math.libretexts.org/@go/page/4167
A graph of S(t) and S (t) is given in Figure 3.4.10. When S (t) < 0 , sales are decreasing; note how at t ≈ 1.16 , S (t) is
′ ′ ′

minimized. That 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 "leveling off."

Figure 3.4.10 : A graph of S(t) in Example 3.4.3 along with S ′


(t) .

Not every critical point corresponds to a relative extrema; f (x) = x has a critical point at (0, 0) but no relative maximum or
3

minimum. Likewise, just because f (x) = 0 we cannot conclude concavity changes at that point. We were 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 without concavity changing is f (x) = x . At x = 0 , f (x) = 0 but f is always concave up, as shown in Figure 3.4.11.
′′ 4 ′′

Figure 3.4.11 : A graph of f (x) = x . Clearly f is always concave up, despite the fact that f (x) = 0 when x = 0 . It this
4 ′′

example, the possible point of inflection (0, 0) is not a point of inflection.

Note: A note about concavity:


In general, concavity can change only where either the second derivative is 0, where there is a vertical asymptote, or (rare in
practice) where the second derivative is undefined. But concavity doesn't \emph{have} to change at these places. For instance,
if f (x) = x , then f (0) = 0 , but there is no change of concavity at 0 and also no inflection point there. Moreover, if
4 ′′

f (x) = 1/x , then f has a vertical asymptote at 0, but there is no change in concavity at 0.
2

The Second Derivative Test


The first derivative of a function gave us a test to find if a critical value corresponded to a relative maximum, minimum, or neither.
The second derivative gives us another way to test if a critical point is a local maximum or minimum. The following theorem
officially states something that is intuitive: if a critical value occurs in a region where a function f is concave up, then that critical
value must correspond to a relative minimum of f , etc. See Figure 3.4.12 for a visualization of this.

3.4.6 https://math.libretexts.org/@go/page/4167
Figure 3.4.12 : Demonstrating the fact that relative maxima occur when the graph is concave down and relatve minima occur
when the graph is concave up.

THeorem 3.4.3: The Second Derivative Test

Let c be a critical value of f where f ′′


(c) is defined.
1. If f ′′
(c) > 0 , then f has a local minimum at (c, f (c)).
2. If f ′′
(c) < 0 , then f has a local maximum at (c, f (c)).

The Second Derivative Test relates to the First Derivative Test in the following way. If f (c) > 0 , then the graph is concave up at a
′′

critical point c and f itself is growing. Since f (c) = 0 and f is growing at c , then it must go from negative to positive at c . This
′ ′ ′

means the function goes from decreasing to increasing, indicating a local minimum at c .

Example 3.4.4: Using the Second Derivative Test

Let f (x) = 100/x + x. Find the critical points of f and use the Second Derivative Test to label them as relative maxima or
minima.
Solution
We find f (x) = −100/x + 1 and f (x) = 200/x . We set f (x) = 0 and solve for x to find the critical values (note that f'\
′ 2 ′′ 3 ′

is not defined at x = 0 , but neither is f so this is not a critical value.) We find the critical values are x = ±10 . Evaluating f ′′

at x = 10 gives 0.1 > 0 , so there is a local minimum at x = 10. Evaluating f (−10) = −0.1 < 0 , determining a relative
′′

maximum at x = −10 . These results are confirmed in Figure 3.4.13.

Figure 3.4.13 : A graph of f (x) in Example 3.4.4. The second derivative is evaluated at each critical point. When the graph
is concave up, the critical point represents a local minimum; when the graph is concave down, the critical point represents a
local maximum.

3.4.7 https://math.libretexts.org/@go/page/4167
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, along with the locations of
relative extrema and inflection points. In Chapter 1 we saw how limits explained asymptotic behavior. In the next section we
combine all of this information to produce accurate sketches of functions.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/
Integrated by Justin Marshall.

This page titled 3.4: Concavity and the Second Derivative is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by
Gregory Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available
upon request.

3.4.8 https://math.libretexts.org/@go/page/4167
3.5: Curve Sketching
We have been learning how we can understand the behavior of a function based on its first and second derivatives. While we have
been treating the properties of a function separately (increasing and decreasing, concave up and concave down, etc.), we combine
them here to produce an accurate graph of the function without plotting lots of extraneous points.
Why bother? Graphing utilities are very accessible, whether on a computer, a hand--held calculator, or a smartphone. These
resources are usually very fast and accurate. We will see that our method is not particularly fast -- it will require time (but it is not
hard). So again: why bother?
We are attempting to understand the behavior of a function f based on the information given by its derivatives. While all of a
function's derivatives relay information about it, it turns out that "most" of the behavior we care about is explained by f and f . ′ ′′

Understanding the interactions between the graph of f and f and f is important. To gain this understanding, one might argue that
′ ′

all that is needed is to look at lots of graphs. This is true to a point, but is somewhat similar to stating that one understands how an
engine works after looking only at pictures. It is true that the basic ideas will be conveyed, but "hands--on'' access increases
understanding.
The following Key Idea summarizes what we have learned so far that is applicable to sketching graphs of functions and gives a
framework for putting that information together. It is followed by several examples.

Key Idea 4: Curve Sketching


To produce an accurate sketch a given function f , consider the following steps.
1. Find the domain of f . Generally, we assume that the domain is the entire real line then find restrictions, such as where a
denominator is 0 or where negatives appear under the radical.
2. Find the critical values of f .
3. Find the possible points of inflection of f .
4. Find the location of any vertical asymptotes of f (usually done in conjunction with item 1 above).
5. Consider the limits lim f (x) and lim f (x) to determine the end behavior of the function.
x→−∞ x→∞

6. Create a number line that includes all critical points, possible points of inflection, and locations of vertical asymptotes. For
each interval created, determine whether f is increasing or decreasing, concave up or down.
7. Evaluate f at each critical point and possible point of inflection. Plot these points on a set of axes. Connect these points
with curves exhibiting the proper concavity. Sketch asymptotes and x and y intercepts where applicable.

Example 3.5.1: curve sketching

Use Key Idea 4 to sketch f (x) = 3x 3


− 10 x
2
+ 7x + 5 .
Solution
1. The domain of f is the entire real line; there are no values x for which f (x) is not defined.
2. Find the critical values of f . We compute f (x) = 9x − 20x + 7 . Use the Quadratic Formula to find the roots of f :
′ 2 ′

−−−−−−−−−−−− −
2
20 ± √ (−20 ) − 4(9)(7) 1 −−
x = = (10 ± √37) ⇒ x ≈ 0.435, 1.787. (3.5.1)
2(9) 9

3. Find the possible points of inflection of f . Compute f ′′


(x) = 18x − 20 . We have

f p(x) = 0 ⇒ x = 10/9 ≈ 1.111. (3.5.2)

4. There are no vertical asymptotes.


5. We determine the end behavior using limits as x approaches ±infinity.

lim f (x) = −∞ lim f (x) = ∞. (3.5.3)


x→−∞ x→∞

We do not have any horizontal asymptotes.


−−
6. We place the values x = (10 ± √37)/9 and x = 10/9 on a number line, as shown in Figure 3.5.1. We mark each
subinterval as increasing or decreasing, concave up or down, using the techniques used in Sections 3.3 and 3.4.

3.5.1 https://math.libretexts.org/@go/page/4168
Figure 3.5.1 : Number line for f in Example 3.5.1.
7. We plot the appropriate points on axes as shown in Figure 3.5.2a and connect the points with straight lines. In Figure
3.5.2b we adjust these lines to demonstrate the proper concavity. Our curve crosses the y axis at y = 5 and crosses the x

axis near x = −0.424. In Figure 3.5.2c we show a graph of f drawn with a computer program, verifying the accuracy of
our sketch.

Figure 3.5.2 : Sketching f in Example 3.5.1.

Example 3.5.2: Curve sketching


2
x −x −2
Sketch f (x) = 2
.
x −x −6

Solution
We again follow the steps outlined in Key Idea 4.
1. In determining the domain, we assume it is all real numbers and looks for restrictions. We find that at x = −2 and x = 3 ,
f (x) is not defined. So the domain of f is D = {real numbers x | x ≠ −2, 3} .

2. To find the critical values of f , we first find f (x). Using the Quotient Rule, we find

−8x + 4 −8x + 4

f (x) = = . (3.5.4)
2 2 2 2
(x + x − 6) (x − 3 ) (x + 2 )


f (x) = 0 when x = 1/2, and f is undefined when x = −2, 3. Since f is undefined only when f is, these are not critical
′ ′

values. The only critical value is x = 1/2.


3. To find the possible points of inflection, we find f (x), again employing the Quotient Rule:
′′

2
24 x − 24x + 56
′′
f (x) = . (3.5.5)
3 3
(x − 3 ) (x + 2 )

We find that f (x) is never 0 (setting the numerator equal to 0 and solving for x, we find the only roots to this quadratic
′′

are imaginary) and f is undefined when x = −2, 3. Thus concavity will possibly only change at x = −2 and x = 3 .

4. The vertical asymptotes of f are at x = −2 and x = 3 , the places where f is undefined.


5. There is a horizontal asymptote of y = 1 , as lim f (x) = 1 and lim
x→−∞ f (x) = 1 . x→∞

6. We place the values x = 1/2, x = −2 and x = 3 on a number line as shown in Figure 3.5.3. We mark in each interval
whether f is increasing or decreasing, concave up or down. We see that f has a relative maximum at x = 1/2; concavity
changes only at the vertical asymptotes.

3.5.2 https://math.libretexts.org/@go/page/4168
FIgure 3.5.3 : Number line for f in Example 3.5.2
7. In Figure 3.5.4a, 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 3.5.4b, we adjust the graph with the appropriate concavity. We also show f crossing the x axis at x = −1 and
x = 2.

Figure 3.5.4 : Sketching f in Example 3.5.2.


Figure 3.5.4c shows a computer generated graph of f , which verifies the accuracy of our sketch.

Example 3.5.3: Curve sketching


5(x−2)(x+1)
Sketch f (x) = x2 +2x+4
.

Solution
We again follow Key Idea 4
1. We assume that the domain of f is all real numbers and consider restrictions. The only restrictions come when the
denominator is 0, but this never occurs. Therefore the domain of f is all real numbers, R.
2. We find the critical values of f by setting f (x) = 0 and solving for x. We find

15x(x + 4)
′ ′
f (x) = ⇒ f (x) = 0 when x = −4, 0. (3.5.6)
2 2
(x + 2x + 4 )

3. We find the possible points of inflection by solving f ′′


(x) = 0 for x. We find
3 2

30 x + 180 x − 240
f p(x) = − . (3.5.7)
2 3
(x + 2x + 4 )

The cubic in the numerator does not factor very "nicely." We instead approximate the roots at x = −5.759, x = −1.305
and x = 1.064.
4. There are no vertical asymptotes.
5. We have a horizontal asymptote of y = 5 , as lim f (x) = lim
x→−∞ f (x) = 5 .
x→∞

6. We place the critical points and possible points on a number line as shown in Figure 3.5.5 and mark each interval as
increasing/decreasing, concave up/down appropriately.

3.5.3 https://math.libretexts.org/@go/page/4168
Figure 3.5.5 : Number line for f in Example 3.5.3.
7. In Figure 3.5.6a we plot the significant points from the number line as well as the two roots of f , x = −1 and x = 2 , and
connect the points with straight lines to get a general impression about the graph. In Figure 3.5.6b, we add concavity.
Figure 3.5.6c shows a computer generated graph of f , affirming our results.

Figure 3.5.6 : Sketching f in Example 3.5.3.

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 connected these points with straight lines, then adjusted for concavity, and finished by showing a very accurate,
computer generated graph.
Why are computer graphics so good? It is not because computers are "smarter" than we are. Rather, it is largely because computers
are much faster at computing 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 connect 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, use special algorithms to plot lots of points only where the graph is
"curvy.''
In Figure 3.5.7, a graph of y = sin x is given, generated by Mathematica. The small points represent each of the places
Mathematica 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. (Many points are also used at the endpoints to ensure the "end behavior" is accurate.)

Figure 3.5.7 : A graph of y = sin x generated by Mathematica.


How does Mathematica know where the graph is "curvy"? Calculus. When we study curvature in a later chapter, we will see how
the first and second derivatives of a function work together to provide a measurement of "curviness." Mathematica employs

3.5.4 https://math.libretexts.org/@go/page/4168
algorithms to determine regions of "high curvature"' and plots extra points there.
Again, the goal of this section is not "How to graph a function when there is no computer to help.'' Rather, the goal is "Understand
that the shape of the graph of a function is largely determined by understanding the behavior of the function at a few key places." In
Example 3.5.3, we were able to accurately sketch a complicated graph using only 5 points and knowledge of asymptotes!
There are many applications of our understanding of derivatives beyond curve sketching. The next chapter explores some of these
applications, demonstrating just a few kinds of problems that can be solved with a basic knowledge of differentiation.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/
Integrated by Justin Marshall.

This page titled 3.5: Curve Sketching is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et
al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

3.5.5 https://math.libretexts.org/@go/page/4168
3.E: Applications of the Graphical Behavior of Functions(Exercises)
3.1: Extreme Values
Terms and Concepts
1. Describe what an "extreme value" of a function is in your own words.
2. Sketch the graph of a function f on (-1,1) that has both a maximum and minimum value.
3. Describe the difference between absolute and relative maxima in your own words.
4. Sketch the graph of a function f where f has a relative maximum at x = 1 and f ′
(1) is undefined.
5. T/F: If c is a critical value of a function f , then f has either a relative maximum or relative minimum at x = c .

Problems
In Exercises 6-7, identify each of the marked points as being an absolute maximum and minimum, a relative maximum or
minimum, or none of the above.
6.

7.

In Exercises 8-14, evaluate f ′


(x) at the points indicated in the graph.
8. f (x) = 2
2
x +1

−−−− −
9. f (x) = x 2
√6 − x2

3.E.1 https://math.libretexts.org/@go/page/9971
10. f (x) = sin x

−−−−−
11. f (x) = x 2
√4 − x

2
x x ≤0
12. f (x) = { 5
x x >0

2
x x ≤0
13. f (x) = {
x x >0

2/3
(x−2)
14. f (x) = x

In Exercises 15-24, find the extreme values of the function on the given interval.
15. f (x) = x 2
+ x + 4 on [−1, 2] .
16. f (x) = x 3

9

2
2
x − 30x + 3 on [0, 6] .

3.E.2 https://math.libretexts.org/@go/page/9971
17. f (x) = 3 sin x on [π/4, 2π/3].
−−−− −
18. f (x) = x 2
√4 − x2 on [−2, 2] .
19. f (x) = x + 3

x
on [1, 5] .
2

20. f (x) = x +5
x
2
on [−3, 5] .
21. f (x) = e x
cos x on [0, π] .
22. f (x) = e x
sin x on [0, π] .
23. f (x) = ln x

x
on [1, 4] .
24. f (x) = x 2/3
− x on [0, 2] .

Review
dy
25. Find dx
, where x 2
y −y x = 1
2
.
26. Find the equation of the line tangent to the graph of x 2
+y
2
+ xy = 7 at the point (1, 2).
f (x+s)−f (x)
27. Let f (x) = x 3
+x . Evaluate lim s
.
s→0

3.2: The Mean Value Theorem


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
In Exercises 3-10, 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].
5. f (x) = x 2
+ x − 6 on [−3, 2] .
6. f (x) = x 2
+ x − 2 on [−3, 2] .
7. f (x) = x 2
+ x on [−2, 2] .
8. f (x) = sin x on [π/6, 5π/6].
9. f (x) = cos x on [0, π].
10. f (x) = 2
x −2x+1
1
on [0, 2] .

In Exercises 11-20, a function f (x) and interval [a,b] are given. Check if the Mean Value Theorem can be applied to f on
[a,b]; if so, find a value c in [a,b] guaranteed by the Mean Value Theorem.
11. f (x) = x 2
+ 3x − 1 on [−2, 2] .
12. f (x) = 5x 2
− 6x + 8 on [0, 5] .
−−−− −
13. f (x) = √9 − x
2
on [0, 3] .
−−−−−
14. f (x) = √25 − x on [0, 9] .
16. f (x) = ln x on [1, 5].
17. f (x) = tan x on [π/4, π/4].
18. f (x) = x 3
− 2x
2
+ x + 1 on [−2, 2] .
19. f (x) = 2x 3
− 5x
2
+ 6x + 1 on [−5, 2] .

3.E.3 https://math.libretexts.org/@go/page/9971
20. f (x) = sin −1
x on [−1, 1] .

Review
21. Find the extreme values of f (x) = x 2
− 3x + 9 on [−2, 5] .
22. Describe the critical points of f (x) = cos x.
23. Describe the critical points of f (x) = tan x.

3.3: Increasing and Decreasing Functions


Terms and Concepts
1. In your own words describe what it means for a function to be increasing.
2. What does a decreasing function “look like”?
3. Sketch a graph of a function on [0,2] that is increasing but not strictly increasing.
4. Give an example of a function describing a situation where it is “bad” to be increasing and “good” to be decreasing.
2

5. A function f has derivative f '(x) = (sin x + 2)e x +1


, where f '(x) > 1 for all x. Is f increasing, decreasing, or can we not tell
from the given information?

Problems
In Exercises 6-13, a function f (x) is given.
(a) Compute f (x). ′

(b) Graph f and f on the same axes (using technology is permitted) and verify Theorem 29.

6. f (x) = 3x + 4
7. f (x) = x 2
− 3x + 5

8. f (x) = cos x
9. f (x) = tan x
10. f (x) = x 3
− 5x
2
+ 7x − 1

11. f (x) = 2x 3
−x
2
+x −1

12. f (x) = x 4
− 5x
2
+4

13. f (x) = 1

x2 +1

In Exercises 14-23, a function f (x) is given.


(a) Give the domain of f .
(b) Find the critical numbers of f .
(c) Create a number line to determine the intervals on which f is increasing and decreasing.
(d) Use the First Derivative Test to determine whether each critical point is a relative maximum, minimum, or neither.
14. f (x) = x 2
+ 2x − 3

15. f (x) = x 3
+ 3x
2
+3

16. f (x) = 2x 3
+x
2
+3

17. f (x) = x 3
− 3x
2
+ 3x − 1

18. f (x) = 2
x −2x+2
1

19. f (x) = x −4

x2 −1

20. f (x) = 2
x −2x−8
x

2/3
(x−2)
21. f (x) = x

22. f (x) = sin x cos x on (−π, π)

3.E.4 https://math.libretexts.org/@go/page/9971
23. f (x) = 5x 2
− 5x

Review
24. Consider f (x) = x 2
− 3x + 5 on [-1,2]; find c guaranteed by the Mean Value Theorem.
25. Consider f (x) = sin x on [−π/2, π/2]; find \(c) guaranteed by the Mean Value Theorem.

3.4: Concavity and the Second Derivative


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:
(a) Increasing, concave up on (0,1),
(b) increasing, concave down on (1,2),
(c) decreasing, concave down on (2,3) and
(d) increasing, concave down on (3,4).
3. Is it possible for a function to be increasing and concave down on (0, ∞) with a horizontal asymptote of y =1 ? If so, give a
sketch of such a function.
4. Is it possible for a function to be increasing and concave upon (0, ∞) with a horizontal asymptote of y = 1 ? If so, give a sketch
of such a function.

Problems
In Exercises 5-15, a function f (x) is given.
(a) Compute f (x). ′′

(b) Graph f and f on the same axes (using technology is permitted) and verify Theorem 31.
′′

5. f (x) = −7x + 3
6. f (x) = −4x 2
+ 3x − 8

7. f (x) = 4x 2
+ 3x − 8

8. f (x) = x 3
− 3x
2
+x −1

9. f (x) = −x 3
+x
2
− 2x + 5

10. f (x) = cos x


11. f (x) = sin x
12. f (x) = tan x
13. f (x) = x +1
1
2

14. f (x) = 1

15. f (x) = 1

x
2

In Exercises 16-28, a function f (x) is given.


(a) Find possible points of inflection of f
(b) Create a number line to determine the intervals on which f is concave up or concave down.
16. f (x) = x 2
− 2x + 1

17. f (x) = −x 2
− 5x + 7

18. f (x) = x 3
−x +1

19. f (x) = 2x 3
− 3x
2
+ 9x + 5

4 3

20. f (x) = x

4
+
x

3
− 2x + 3

21. f (x) = −3x 4


+ 8x
3
+ 6x
2
− 24x + 2

3.E.5 https://math.libretexts.org/@go/page/9971
22. f (x) = x 4
− 4x
3
+ 6x
2
− 4x + 1

23. f (x) = 2
x +1
1

24. f (x) = 2
x −1
x

25. f (x) = sin x + cos x on (−π, π)


26. f (x) = x 2
e
x

27. f (x) = x 2
ln x

28. f (x) = e
2
−x

In Exercises 29-41, 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 exercises 16-28.)
29. f (x) = x 2
− 2x + 1

30. f (x) = −x 2
− 5x + 7

31. f (x) = x 3
−x +1

32. f (x) = 2x 3
− 3x
2
+ 9x + 5

4 3

33. f (x) = x

4
+
x

3
− 2x + 3

34. f (x) = −3x 4


+ 8x
3
+ 6x
2
− 24x + 2

35. f (x) = x 4
− 4x
3
+ 6x
2
− 4x + 1

36. f (x) = 1

x2 +1

37. f (x) = x

x2 −1

38. f (x) = sin x + cos x on (−π, π)


39. f (x) = x 2
e
x

40. f (x) = x 2
ln x

41. f (x) = e
2
−x

In Exercises 42-54, 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 Exercises 16-28.)
42. f (x) = x 2
− 2x + 1

43. f (x) = −x 2
− 5x + 7

44. f (x) = x 3
−x +1

45. f (x) = 2x 3
− 3x
2
+ 9x + 5

4 3

46. f (x) = x

4
+
x

3
− 2x + 3

47. f (x) = −3x 4


+ 8x
3
+ 6x
2
− 24x + 2

48. f (x) = x 4
− 4x
3
+ 6x
2
− 4x + 1

49. f (x) = 2
x +1
1

50. f (x) = 2
x −1
x

51. f (x) = sin x + cos x on (−π, π)


52. f (x) = x 2
e
x

53. f (x) = x 2
ln x

54. f (x) = e
2
−x

3.E.6 https://math.libretexts.org/@go/page/9971
3.5: Curve Sketching
Terms and Concepts
1. Why is sketching curves by hand beneficial even though technology is ubiquitous?
2. What does “ubiquitous” mean?
3. T/F: When sketching graphs of functions, it is useful to find the critical points.
4. T/F: When sketching graphs of functions, it is useful to find the possible points of inflection.
5. T/F: When sketching graphs of functions, it is useful to find the horizontal and vertical asymptotes.

Problems
In Exercises 6-11, practice using Key Idea 4 by applying the principles to the given functions with familiar graphs.
6. f (x) = 2x + 4
7. f (x) = −x 2
+1

8. f (x) = sin x
9. f (x) = e x

10. f (x) = 1

11. f (x) = 1

x
2

In Exercises 12-25, sketch a graph of the given function using Key Idea 4. Show all work; check your answer with
technology.
12. f (x) = x 3
− 2x
2
+ 4x + 1

13. f (x) = −x 3
+ 5x
2
− 3x + 2

14. f (x) = x 3
+ 3x
2
+ 3x + 1

15. f (x) = x 3
−x
2
−x +1

16. f (x) = (x − 2) ln(x − 2)


17. f (x) = (x − 2) 2
ln(x − 2)

2
x −4
18. f (x) = x
2

2
x −4x+3
19. f (x) = 2
x −6x+8

20. f (x) = x −2x+1

x2 −6x+8

−−−−−
21. f (x) = x √x + 1
22. f (x) = x 2
e
x

23. f (x) = sin x cos x on [−π, π]


24. f (x) = (x − 3) 2/3
+2

2/3
(x−1)
25. f (x) = x

In Exercises 26-28, 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 .
26. f (x) = 2
a
2
x +b

27. f (x) = sin(ax + b)


28. f (x) = (x − a)(x − b)

3.E.7 https://math.libretexts.org/@go/page/9971
2
d y
29. Given x 2
+y
2
=1 , use implicit differentiation to find f racdydx and dx
2
. Use this information to justify the sketch of the unit
circle.

3.E: Applications of the Graphical Behavior of Functions(Exercises) is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or
curated by LibreTexts.

3.E.8 https://math.libretexts.org/@go/page/9971
CHAPTER OVERVIEW
4: Applications of the Derivative
4.1: Newton's Method
4.2: Related Rates
4.3: Optimization
4.4: Differentials
4.E: Applications of Derivatives (Exercises)

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 4: Applications of the Derivative is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory
Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

1
4.1: Newton's Method
In Chapter 3, we learned how the first and second derivatives of a function influence its graph. In this chapter we explore other
applications of the derivative.
Solving equations is one of the most important things we do in mathematics, yet we are surprisingly limited in what we can solve
analytically. For instance, equations as simple as x + x + 1 = 0 or cos x = x cannot be solved by algebraic methods in terms of
5

familiar functions. Fortunately, there are methods that can give us approximate solutions to equations like these. These methods can
usually give an approximation correct to as many decimal places as we like. In Section 1.5, we learned about the Bisection Method.
This section focuses on another technique (which generally works faster), called Newton's Method.
Newton's Method is built around tangent lines. The main idea is that if x is sufficiently close to a root of f (x), then the tangent line
to the graph at (x, f (x)) will cross the x-axis at a point closer to the root than x.

Figure 4.1.1 : Demonstrating the geometric concept behind Newton's Method


We start Newton's Method with an initial guess about roughly where the root is. Call this x . (See Figure 4.1.1a.) Draw the tangent
0

line to the graph at (x , f (x )) and see where it meets the x-axis. Call this point x . Then repeat the process -- draw the tangent
0 0 1

line to the graph at (x , f (x )) and see where it meets the x-axis. (See Figure 4.1.1b.) Call this point x . Repeat the process again
1 1 2

to get x , x , etc. This sequence of points will often converge rather quickly to a root of f .
3 4

We can use this geometric process to create an algebraic process. Let's look at how we found x . We started with the tangent line
1

to the graph at (x , f (x )). The slope of this tangent line is f (x ) and the equation of the line is
0 0

0

$$y=f'(x_0)(x-x_0)+f(x_0).\]
This line crosses the x-axis when y = 0 , and the x--value where it crosses is what we called x . So let 1 y =0 and replace x with
x , giving the equation:
1


0 = f (x0 )(x1 − x0 ) + f (x0 ). (4.1.1)

Now solve for x :1

f (x0 )
x1 = x0 − . (4.1.2)

f (x0 )

Since we repeat the same geometric proces to find x from x , we have


2 1

f (x1 )
x2 = x1 − . (4.1.3)

f (x1 )

In general, given an approximation x , we can find the next approximation, x


n n+1 as follows:
f (xn )
xn+1 = xn − . (4.1.4)

f (xn )

We summarize this process as follows.

4.1.1 https://math.libretexts.org/@go/page/4171
Key Idea 5: Newton's Method
Let f be a differentiable function on an interval I with a root in I . To approximate the value of the root, accurate to d decimal
places:
1. Choose a value x as an initial approximation of the root. (This is often done by looking at a graph of f .)
0

2. Create successive approximations iteratively; given an approximation x , compute the next approximation x
n n+1 as
f (xn )
xn+1 = xn − . (4.1.5)

f (xn )

3. Stop the iterations when successive approximations do not differ in the first d places after the decimal point.

Note: Newton's Method is not infallible. The sequence of approximate values may not converge, or it may converge so slowly that
one is "tricked" into thinking a certain approximation is better than it actually is. These issues will be discussed at the end of the
section.}
Let's practice Newton's Method with a concrete example.

Example 4.1.1: Using Newton's Method

Approximate the real root of x − x 3 2


−1 = 0 , accurate to the first 3 places after the decimal, using Newton's Method and an
initial approximation of x = 1 .}
0

Solution
To begin, we compute f ′
(x) = 3 x
2
− 2x . Then we apply the Newton's Method algorithm, outlined in Key Idea 5.
3 2
f (1) 1 −1 −1
x1 =1− =1− = 2, (4.1.6)
′ 2
f (1) 3⋅1 −2 ⋅ 1
3 2
f (2) 2 −2 −1
x2 =2− =2− = 1.625, (4.1.7)
′ 2
f (2) 3⋅2 −2 ⋅ 2
3 2
f (1.625) 1.625 − 1.625 − 1
x3 = 1.625 − = 1.625 − ≈ 1.48579. (4.1.8)
′ 2
f (1.625) 3 ⋅ 1.625 − 2 ⋅ 1.625

f (1.48579)
x4 = 1.48579 − ≈ 1.46596 (4.1.9)

f (1.48579)

f (1.46596)
x5 = 1.46596 − ≈ 1.46557 (4.1.10)

f (1.46596)

We performed 5 iterations of Newton's Method to find a root accurate to the first 3 places after the decimal; our final
approximation is 1.465. The exact value of the root, to six decimal places, is 1.465571; It turns out that our x is accurate to 5

more than just 3 decimal places.


A graph of f (x) is given in Figure 4.1.2. We can see from the graph that our initial approximation of x = 1 was not 0

particularly accurate; a closer guess would have been x = 1.5 . Our choice was based on ease of initial calculation, and shows
0

that Newton's Method can be robust enough that we do not have to make a very accurate initial approximation.

4.1.2 https://math.libretexts.org/@go/page/4171
Figure 4.1.2 : A graph of f (x) = x 3
−x
2
−1 in Example 4.1.1.

We can automate this process on a calculator that has an Ans key that returns the result of the previous calculation. Start by pressing
1 and then Enter . (We have just entered our initial guess, x = 1 .) Now compute
0

$$\texttt{Ans} - \frac{f(\texttt{Ans})}{f'(\texttt{Ans})}\]
by entering the following and repeatedly press the Enter key:
3 2 2
Ans − (Ans − Ans − 1)/(3 ∗ Ans − 2 ∗ Ans) (4.1.11)

Each time we press the \texttt{Enter} key, we are finding the successive approximations, x , x , \dots, and each one is getting
1 2

closer to the root. In fact, once we get past around x or so, the approximations don't appear to be changing. They actually are
7

changing, but the change is far enough to the right of the decimal point that it doesn't show up on the calculator's display. When this
happens, we can be pretty confident that we have found an accurate approximation.
Using a calculator in this manner makes the calculations simple; many iterations can be computed very quickly.

Example 4.1.2: Using Newton's Method to find where functions intersect

Use Newton's Method to approximate a solution to cos x = x, accurate to 5 places after the decimal.
Solution
Newton's Method provides a method of solving f (x) = 0 ; it is not (directly) a method for solving equations like f (x) = g(x) .
However, this is not a problem; we can rewrite the latter equation as f (x) − g(x) = 0 and then use Newton's Method.
So we rewrite cos x = x as cos x − x = 0 . Written this way, we are finding a root of f (x) = cos x − x . We compute

f (x) = − sin x − 1 . Next we need a starting value, x . Consider Figure 4.1.3, where f (x) = cos x − x is graphed. It seems
0

that x = 0.75 is pretty close to the root, so we will use that as our x . (The figure also shows the graphs of y = cos x and
0 0

y = x , drawn with dashed lines. Note how they intersect at the same x value as when f (x) = 0 .)

Figure 4.1.3 : A graph of f (x) = cos x − x used to find an initial approximation of its root.
We now compute x , x , etc. The formula for x is
1 2 1

cos(0.75) − 0.75
x1 = 0.75 − ≈ 0.7391111388. (4.1.12)
− sin(0.75) − 1

Apply Newton's Method again to find x : 2

cos(0.7391111388) − 0.7391111388
x2 = 0.7391111388 − ≈ 0.7390851334. (4.1.13)
− sin(0.7391111388) − 1

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
pressing 0.75, then Enter , inputting our initial approximation. We then enter:

Ans - (cos(Ans)-Ans)/(-sin(Ans)-1). (4.1.14)

4.1.3 https://math.libretexts.org/@go/page/4171
Repeatedly pressing the Enter key gives successive approximations. We quickly find:
x3 = 0.7390851332 (4.1.15)

x4 = 0.7390851332. (4.1.16)

Our approximations x and x did not differ for at least the first 5 places after the decimal, so we could have stopped.
2 3

However, using our calculator in the manner described is easy, so finding x was not hard. It is interesting to see how we found
4

an approximation, accurate to as many decimal places as our calculator displays, in just 4 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 = .75

while true

oldx = x

x = x - (cos(x)-x)/(-sin(x)-1)

print x

if abs(x-oldx) < .0000000001

break

This code calculates x , x , etc., storing each result in the variable x . The previous approximation is stored in the variable oldx .
1 2

We continue looping until the difference between two successive approximations, abs(x-oldx) , is less than some small
tolerance, in this case, .0000000001 .

Convergence of Newton's Method


What should one use for the initial guess, x ? Generally, the closer to the actual root the initial guess is, the better. However, some
0

initial guesses should be avoided. For instance, consider Example 4.1.1 where we sought the root to f (x) = x − x − 1 . 3 2

Choosing x = 0 would have been a particularly poor choice. Consider Figure 4.1.4, where f (x) is graphed along with its tangent
0

line at x = 0 . Since f (0) = 0 , the tangent line is horizontal and does not intersect the x--axis. Graphically, we see that Newton's

Method fails.

\Figure 4.1.4 : A graph of f (x) = x 3


−x
2
−1 , showing why an initial approximation of x 0 =0 with Newton's Method fails.
We can also see analytically that it fails. Since
f (0)
x1 = 0 − (4.1.17)

f (0)

and f ′
(0) = 0 , we see that x is not well defined.
1

This problem can also occur if, for instance, it turns out that ′
f (x5 ) = 0 . Adjusting the initial approximation x0 by a very small
amount will likely fix the problem.

4.1.4 https://math.libretexts.org/@go/page/4171
It is also possible for Newton's Method to not converge while each successive approximation is well defined. Consider
f (x) = x
1/3
, as shown in Figure 4.1.5. It is clear that the root is x = 0 , but let's approximate this with x = 0.1 . Figure 4.1.5a
0

shows graphically the calculation of x ; notice how it is farther from the root than x . Figures 4.1.5b and (4.1.5c) show the
1 0

calculation of x and x , which are even farther away; our successive approximations are getting worse. (It turns out that in this
2 3

particular example, each successive approximation is twice as far from the true answer as the previous approximation.)

Figure 4.1.5 : Newton's Method fails to find a root of f (x) = x 1/3


, regardless of the choice of x .}\label{fig:newt4}
0

There is no "fix" to this problem; Newton's Method simply will not work and another method must be used.
While Newton's Method does not always work, it does work "most of the time," and it is generally very fast. Once the
approximations get close to the root, Newton's Method can as much as double the number of correct decimal places with each
successive approximation. A course in Numerical Analysis will introduce the reader to more iterative root finding methods, as well
as give greater detail about the strengths and weaknesses of Newton's Method.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/
Integrated by Justin Marshall.

This page titled 4.1: Newton's Method is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et
al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

4.1.5 https://math.libretexts.org/@go/page/4171
4.2: Related Rates
When two quantities are related by an equation, knowing the value of one quantity can determine the value of the other. For
instance, the circumference and radius of a circle are related by C = 2πr; knowing that C = 6π in determines the radius must be 3
in.
The topic of related rates takes this one step further: knowing the rate at which one quantity is changing can determine the rate at
which the other changes.
Note: This section relies heavily on implicit differentiation, so referring back to Section 2.6 may help. We demonstrate the concepts
of related rates through examples.

Example 4.2.1: Understanding related rates

The radius of a circle is growing at a rate of 5 in/hr. At what rate is the 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 respect to time; that is, we are told = 5 in/hr. We want to know how the length of C changes with respect to time, i.e.,
dr

dt

we want to know .dC

dt

Implicitly differentiate both sides of C = 2πr with respect to t :

C = 2πr

d d
(C ) = (2πr)
dt dt
dC dr
= 2π .
dt dt

As we know dr

dt
=5 in/hr, we know
$$\frac{dC}{dt} = 2\pi 5 = 10\pi \approx 31.4\text{in/hr.}\]

Consider another, similar example.

Example 4.2.2: Finding related rates

Water streams out of a faucet at a rate of 2in /s onto a flat surface at a constant rate, forming a circular puddle that is
3
1/8 in
deep.
1. At what rate is the area of the puddle growing?
2. At what rate is the radius of the circle growing?
Solution
1. We can answer this question two ways: using "common sense" or related rates. The common sense method states that the
volume of the puddle is growing by 2in /s, where
3

volume of puddle = area of circle × depth. (4.2.1)

Since the depth is constant at 1/8in, the area must be growing by 16in /s. 2

This approach reveals the underlying related--rates principle. Let V and A represent the Volume and Area of the puddle. We
know V = A × . Take the derivative of both sides with respect to t , employing implicit differentiation.
1

1
V = A (4.2.2)
8

d d 1
(V ) = ( A) (4.2.3)
dt dt 8

dV 1 dA
= (4.2.4)
dt 8 dt

4.2.1 https://math.libretexts.org/@go/page/4172
As dV

dt
=2 , we know 2 = 1

8
dA

dt
, and hence dA

dt
= 16 . Thus the area is growing by 16in /s.
2

2. To start, we need an equation that relates what we know to the radius. We just learned something about the surface area of
the circular puddle, and we know A = πr . We should be able to learn about the rate at which the radius is growing with this
2

information.
Implicitly derive both sides of A = πr with respect to t :
2

2
A = πr (4.2.5)

d d
2
(A) = (π r ) (4.2.6)
dt dt

dA dr
= 2πr (4.2.7)
dt dt

Our work above told us that dA

dt
= 16 in
2
/s. Solving for dr

dt
, we have
$$\frac{dr}{dt} = \frac{8}{\pi r}.\]
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 how big the circle already is. If the circle is very large, adding 2in of water will not make the circle much bigger
3

at all. If the circle 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 radius in order to know a rate of
change. When the puddle has a radius of 10in, the radius is growing at a rate of
$$\frac{dr}{dt} = \frac{8}{10\pi} = \frac{4}{5\pi} \approx 0.25\text{in/s}.\]

Example 4.2.3: Studying related rates

Radar guns measure the rate of distance change between the gun and the object it is measuring. For instance, a reading of
"55mph" means the object is moving away from the gun at a rate of 55 miles per hour, whereas a measurement of "−25mph"
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 readouts are only immediately understandable if the gun
and the object are moving along the same line. If a police officer is traveling 60mph and gets a readout of 15mph, he knows
that the car ahead of him is moving away at a rate of 15 miles an hour, meaning the car is traveling 75mph. (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
accurate reading.)
Suppose an officer is driving due north at 30 mph and sees a car moving due east, as shown in Figure 4.2.1. Using his radar
gun, he measures a reading of 20mph. By using landmarks, he believes both he and the other car are about 1/2 mile from the
intersection of their two roads.

Figure 4.2.1 : A sketch of a police car (at bottom) attempting to measure the speed of a car (at right) in Example 4.2.3.
If the speed limit on the other road is 55mph, is the other driver speeding?

4.2.2 https://math.libretexts.org/@go/page/4172
Solution
Using the diagram in Figure 4.2.1, 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.
We know the police officer is traveling at 30mph; that is, = −30 . The reason this rate of change is negative is that A is
dA

dt

getting smaller; the distance between the officer and the intersection is shrinking. The radar measurement is = 20 . We
dC

dt

want to find dB

dt
.
We need an equation that relates B to A and/or C . The Pythagorean Theorem is a good choice: A 2
+B
2
=C
2
. Differentiate
both sides with respect to t :
2 2 2
A +B =C (4.2.8)

d 2 2
d 2
(A +B ) = (C ) (4.2.9)
dt dt

dA dB dC
2A + 2B = 2C (4.2.10)
dt dt dt

We have values for everything except dB

dt
. Solving for this we have
$$\frac{dB}{dt} = \frac{C\frac{dC}{dt}- A\frac{dA}{dt}}{B} \approx 58.28\text{mph}.\]
The other driver appears to be speeding slightly.

Note: Example 4.2.3 is both interesting and impractical. It highlights the difficulty in using radar in a non--linear fashion, and
explains why "in real life" the police officer would follow the other driver to determine their speed, and not pull out pencil and
paper.
The principles here are important, though. Many automated vehicles make judgments about other moving objects based on
perceived distances, radar--like measurements and the concepts of related rates.

Example 4.2.4: Studying related rates

A camera is placed on a tripod 10ft from the side of a road. The camera is to turn to track a car that is to drive by at 100mph for
a promotional video. The video's planners want to know what kind of motor the tripod should be equipped with in order to
properly track the car as it passes by. Figure 4.2.2 shows the proposed setup.

Figure 4.2.2 : Tracking a speeding car (at left) with a rotating camera.
How fast must the camera be able to turn to track the car?
Solution
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 car.
Figure 4.2.2 suggests we use a trigonometric equation. Letting x represent the distance the car is from the point on the road
directly in front of the camera, we have
x
tan θ = (4.2.11)
10

4.2.3 https://math.libretexts.org/@go/page/4172
As the car is moving at 100mph, we have = −100 mph (as in the last example, since x is getting smaller as the car travels,
dx

dt
dx

dt
is negative). We need to convert the measurements so they use the same units; rewrite -100mph in terms of ft/s:
$$\frac{dx}{dt} = -100\frac{\text{m}}{\text{hr}} = -100\frac{\text{m}}{\text{hr}}\cdot5280\frac{\text{ft}}
{\text{m}}\cdot\frac{1}{3600}\frac{\text{hr}}{\text{s}} =-146.\overline{6}\text{ft/s}.\]
Now take the derivative of both sides of Equation 4.2.9 using implicit differentiation:
x
tan θ = (4.2.12)
10
d d x
( tan θ) = ( ) (4.2.13)
dt dt 10

2
dθ 1 dx
sec θ = (4.2.14)
dt 10 dt
2
dθ cos θ dx
= (4.2.15)
dt 10 dt

We want to know the fastest the camera has to turn. Common sense tells us this is when the car is directly in front of the
camera (i.e., when θ = 0 ). Our mathematics bears this out. In Equation 4.2.14 we see this is when cos θ is largest; this is
2

when cos θ = 1 , or when θ = 0 .


With dx

dt
≈ −146.67 ft/s, we have
$$\frac{d\theta}{dt} = -\frac{1\text{rad}}{10\text{ft}}146.67\text{ft/s} = -14.667\text{radians/s}.\]
We find that dθ

dt
is negative; this matches our diagram in Figure 4.2.2 for θ is getting smaller as the car approaches the camera.
What is the practical meaning of −14.667 radians/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.

We introduced the derivative as a function that gives the slopes of tangent lines of functions. This chapter emphasizes using the
derivative in other ways. Newton's Method uses the derivative to approximate roots of functions; this section stresses the "rate of
change" aspect of the derivative to find a relationship between the rates of change of two related quantities. In the next section we
use Extreme Value concepts to optimize quantities.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/
Integrated by Justin Marshall.

This page titled 4.2: Related Rates is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et al.
via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

4.2.4 https://math.libretexts.org/@go/page/4172
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 the concepts of extreme values to solve "word problems," i.e.,
problems stated in terms of situations that require us to create the appropriate mathematical framework in which to solve the
problem.
We start with a classic example which is followed by a discussion of the topic of optimization.

Example 4.3.1: Optimization: perimeter and area

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 4.3.1, either with green grass and nice
fence boards or as a simple rectangle. Either way, drawing a rectangle forces us to realize that we need to know the dimensions
of this rectangle so we can create an area function -- after all, we are trying to maximize the area.

Figure 4.3.1 : A sketch of the enclosure in Example 4.3.1.

We let x and y denote the lengths of the sides of the rectangle. Clearly,
Area = xy. (4.3.1)

We do not yet know how to handle functions with 2 variables; we need to reduce this down to a single variable. We know more
about the situation: the man has 100 feet of fencing. By knowing the perimeter of the rectangle must be 100, we can create another
equation:
Perimeter = 100 = 2x + 2y. (4.3.2)

We now have 2 equations and 2 unknowns. In the latter equation, we solve for y :

y = 50 − x. (4.3.3)

Now substitute this expression for y in the area equation:

Area = A(x) = x(50 − x). (4.3.4)

4.3.1 https://math.libretexts.org/@go/page/4173
Note we now have an equation of one variable; we can truly call the Area a function of x.
This function only makes sense when 0 ≤ x ≤ 50 , otherwise we get negative values of area. So we find the extreme values of
A(x) on the interval [0, 50].

To find the critical points, we take the derivative of A(x) and set it equal to 0, then solve for x.
A(x) = x(50 − x) (4.3.5)

2
= 50x − x (4.3.6)

A (x) = 50 − 2x (4.3.7)

We solve 50 − 2x = 0 to find x = 25; this is the only critical point. We evaluate A(x) at the endpoints of our interval and at this
critical point to find the extreme values; in this case, all we care about is the maximum.
Clearly A(0) = 0 and A(50) = 0 , whereas A(25) = 625ft . This is the maximum. Since we earlier found y = 50 − x , we find
2

that y is also 25. Thus the dimensions of the rectangular enclosure with perimeter of 100 ft. with maximum area is a square, with
sides of length 25 ft.
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 describe 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 principles here will provide a good
foundation for the mathematics you will likely encounter later.
We outline here the basic process of solving these optimization problems.

Key Idea 6: Solving Optimization Problems


1. Understand the problem. Clearly identify what quantity is to be maximized or minimized. Make a sketch if helpful.
2. Create equations relevant to the context of the problem, using the information given. (One of these should describe the
quantity to be optimized. We'll call this the fundamental equation.)
3. If the fundamental equation defines the quantity to be optimized as a function of more than one variable, reduce it to a
single variable function using substitutions derived from the other equations.
4. Identify the domain of this function, keeping in mind the context of the problem.
5. Find the extreme values of this function on the determined domain.
6. Identify the values of all relevant quantities of the problem.

We will use Key Idea 6 in a variety of examples.

Example 4.3.2: {Optimization: perimeter and area

Here is another classic calculus problem: A woman has a 100 feet of fencing, a small dog, and a large yard that contains a
stream (that is mostly straight). She wants to create a rectangular enclosure with maximal area that uses the stream as one side.
(Apparently her dog won't swim away.) What dimensions provide the maximal area?
Solution
We will follow the steps outlined by Key Idea 6.
1. We are maximizing area. A sketch of the region will help; Figure 4.3.2 gives two sketches of the proposed enclosed area. A
key feature of the sketches is to acknowledge that one side is not fenced.

4.3.2 https://math.libretexts.org/@go/page/4173
Figure 4.3.2 : A sketch of the enclosure in Example 4.3.2.
2. We want to maximize the area; as in the example before,
Area = xy. (4.3.8)

This is our fundamental equation. This defines area as a function of two variables, so we need another equation to reduce it
to one variable.
We again appeal to the perimeter; here the perimeter is
Perimeter = 100 = x + 2y. (4.3.9)

Note how this is different than in our previous example.


3. We now reduce the fundamental equation to a single variable. In the perimeter equation, solve for y : y = 50 − x/2 . We can
now write Area as
1
2
Area = A(x) = x(50 − x/2) = 50x − x . (4.3.10)
2

Area is now defined as a function of one variable.


4. We want the area to be nonnegative. Since A(x) = x(50 − x/2), we want x ≥ 0 and 50 − x/2 ≥ 0 . The latter inequality
implies that x ≤ 100, so 0 ≤ x ≤ 100 .
5. We now find the extreme values. At the endpoints, the minimum is found, giving an area of 0.
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(50) = 50(25) = 1250. (4.3.11)

6. We earlier set y = 50 − x/2 ; thus y = 25. Thus our rectangle will have two sides of length 25 and one side of length 50,
with a total area of 1250 ft .
2

Keep in mind as we do these problems that we are practicing a process; that is, we are learning to turn a situation into a system of
equations. These equations allow us to write a certain quantity as a function of one variable, which we then optimize.

Example 4.3.3: Optimization: minimizing cost

A power line needs to be run from an power station located on the beach to an offshore facility. Figure 4.3.3 shows the
distances between the power station to the facility.
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 run along the land to minimize the overall cost? What is the minimal cost?

4.3.3 https://math.libretexts.org/@go/page/4173
Figure 4.3.3 : Running a power line from the power station to an offshore facility with minimal cost in Example 4.3.3
Solution
We will follow the strategy of Key Idea 6 implicitly, without 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 costly option. Second, we could minimize the underwater length by running 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 the distances as shown in Figure
4.3.4.

Figure 4.3.4 : Labeling unknown distances in Example 4.3.3.


By choosing x as we did, we make the expression under the square root simple. We now create the cost function.
$$

Cost = land cost + water cost

$50× land distance + $130× water distance (4.3.12)


−−−−−−−−−
2 2
50(5000 − x) + 130 √ x + 1000 .

\]
−−−−−−−−−
So we have c(x) = 50(5000 − x) + 130√x + 1000 . This function only makes sense on the interval
2 2
[0, 5000] . While we
are fairly certain the endpoints will not give a minimal cost, we still evaluate c(x) at each to verify.
$$c(0) = 380,000 \quad\quad c(5000) \approx 662,873.\]
We now find the critical values of c(x). We compute c (x) as

$$c'(x) = -50+\frac{130x}{\sqrt{x^2+1000^2}}.\]
Recognize that this is never undefined. Setting c (x) = 0 and solving for x, we have:

4.3.4 https://math.libretexts.org/@go/page/4173
130x
−50 + − −−−− −−−− =0 (4.3.13)
√ x2 + 10002

130x
= 50 (4.3.14)
− −−−− −−−−
√ x2 + 10002

2 2
130 x 2
= 50 (4.3.15)
2 2
x + 1000
2 2 2 2 2
130 x = 50 (x + 1000 ) (4.3.16)
2 2 2 2 2 2
130 x − 50 x = 50 ⋅ 1000 (4.3.17)
2 2 2 2
(130 − 50 )x = 50, 000 (4.3.18)
2
2
50, 000
x = (4.3.19)
2 2
130 − 50

50, 000
x = − −−−−−− −− (4.3.20)
√ 1302 − 502

50, 000 2
x = = 416 ≈ 416.67. (4.3.21)
120 3

Evaluating c(x) at x = 416.67 gives a cost of about $370,000. The distance the power line is laid along land is
−−−−−−−−−−−−−
5000 − 416.67 = 4583.33 ft., and the underwater distance is √416.67 2
+ 1000
2
≈ 1083 ft.

In the exercises you will see a variety of situations that require you to combine problem--solving skills with calculus. Focus on the
process; learn how to form equations from situations that can be manipulated into what you need. Eschew memorizing how to do
"this kind of problem" as opposed to "that kind of problem." Learning a process will benefit one far longer than memorizing a
specific technique.
The next section 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.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/
Integrated by Justin Marshall.

This page titled 4.3: Optimization is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et al.
via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

4.3.5 https://math.libretexts.org/@go/page/4173
4.4: Differentials
In Section 2.2 we explored the meaning and use of the derivative. This section 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 equation
$$y = f'(c)(x-c)+f(c).\]
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 cos(π/3) = 1/2. Thus the tangent line to f (x) = sin x at x = π/3 is:

$$ \ell(x) = \frac12(x-\pi/3)+0.866.\]

Figure 4.4.1 : Graphing f (x) = sin x and its tangent line at x = π/3 in order to estimate sin 1.1.
In Figure 4.4.1a, 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 4.4.1b. 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 (4.4.1)
2
1
= (0.053) + 0.866 = 0.8925. (4.4.2)
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 ℓ(x) = f ′
(c)(x − c) + f (c) . Clearly,
f (c) = ℓ(c) . Let Δx be a small number, representing a small change in x value. We assert that:

$$f(c+\Delta x) \approx \ell(c+\Delta x),\]


since the tangent line to a function approximates well the values of that function near x = c .
As the x value changes from c to c + Δx , the y value of f changes from f (c) to f (c + Δx) . We call this change of y value Δy .
That is:
$$\Delta y = f(c+\Delta x)-f(c).\]
Replacing f (c + Δx) with its tangent line approximation, we have

4.4.1 https://math.libretexts.org/@go/page/4174
Δy ≈ ℓ(c + Δx) − f (c)


= f (c)((c + Δx) − c) + f (c) − f (c)


= f (c)Δx (4.4.3)

This final equation is important; we'll come back to it in Key Idea 7.


We introduce two new variables, dx and dy in the context of a formal definition.

Definition: Differentials of x and y .

Let y = f (x) be differentiable. The differential of x, denoted dx , is any nonzero real number (usually taken to be a small
number). The differential of y , denoted dy , is

dy = f (x)dx. (4.4.4)

We can solve for ′


f (x) in the above equation: ′
f (x) = dy/dx . This states that the derivative of f with respect to x is the
dy
differential of y divided by the differential of x; this is not the alternate notation for the derivative, . This latter notation was
dx

chosen because of the fraction--like qualities of the derivative, but again, it is one symbol and not a fraction.
It is helpful to organize our new concepts and notations in one place.

Key Idea 7: Differential Notation


Let y = f (x) be a differentiable function.
1. Δx represents a small, nonzero change in x value.
2. dx represents a small, nonzero change in x value (i.e., Δx = dx ).
3. Δy is the change in y value as x changes by Δx; hence

Δy = f (x + Δx) − f (x). (4.4.5)

4. dy = f ′
(x)dx which, by Equation 4.4.7, is an approximation of the change in y value as x changes by Δx; dy ≈ Δy .

What is the value of differentials? Like many mathematical concepts, differentials provide both practical and theoretical benefits.
We explore both here.

Example 4.4.1: Finding and using differentials

Consider f (x) = x . Knowing f (3) = 9 , approximate f (3.1).


2

Solution
The x value is changing from x = 3 to x = 3.1; therefore, we see that dx = 0.1. If we know how much the y value changes
from f (3) to f (3.1) (i.e., if we know Δy), we will know exactly what f (3.1) is (since we already know f (3)). We can
approximate \Delta y\ with dy .

Δy ≈ dy (4.4.6)

= f (3)dx (4.4.7)

= 2 ⋅ 3 ⋅ 0.1 = 0.6. (4.4.8)

We expect the y value to change by about 0.6, so we approximate f (3.1) ≈ 9.6.


We leave it to the reader to verify this, but the preceding discussion links the differential to the tangent line of f (x) at x = 3 .
One can verify that the tangent line, evaluated at x = 3.1, also gives y = 9.6 .

Of course, it is easy to compute the actual answer (by hand or with a calculator): 3.1 2
= 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.

4.4.2 https://math.libretexts.org/@go/page/4174
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. However, we can approximate the path of an object using differentials. Over small
intervals, the path taken by a floating object is essentially linear. Differentials allow us to approximate the true path by piecing
together lots of short, linear paths. This technique is called Euler's Method, studied in introductory Differential Equations courses.
We use differentials once more to approximate the value of a function. Even though calculators are very accessible, it is neat to see
how these techniques can sometimes be used to easily compute something that looks rather hard.

Example 4.4.2: Using differentials to approximate a function value



−−
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 differentials, with dx = 0.5:
$$f(4.5)-f(4) = \Delta y \approx dy = f'(4)\cdot dx = 1/4 \cdot 1/2 = 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.

Differentials are important when we discuss integration. When we study that topic, we will use notation such as
$$\int f(x)\ dx\]
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.

Example 4.4.3: Finding differentials

In each of the following, find the differential dy .


−−−−−−−− −
x 2 2
y = sin x 2.y = e (x + 2) 3.y = √ x + 3x − 1 (4.4.9)

Solution
1. y = sin x : As f (x) = sin x , f ′
(x) = cos x . Thus
dy = cos(x)dx. (4.4.10)

2. y = e (x + 2) : Let f (x) = e (x + 2) . We need f


x 2 x 2 ′
(x) , requiring the Product Rule.
We have f (x) = e (x + 2) + 2x e , so
′ x 2 x

x 2 x
dy = (e (x + 2) + 2x e )dx. (4.4.11)

−−−−−−−− − −−−−−−−− −
3. 2
y = √x + 3x − 1 : Let 2
f (x) = √x + 3x − 1 ; we need f ′
, requiring the Chain Rule.
(x)
1

We have Δsf ′
(x) =
1

2
(x
2
+ 3x − 1 )

2 (2x + 3) =
2x+3

2
. Thus
2 √x +3x−1

(2x + 3)dx
dy = . (4.4.12)
−−−−−−−− −
2 √ x2 + 3x − 1

Finding the differential dy of y = f (x) is really no harder than finding the derivative of f ; we just multiply ′
f (x) by dx . It is
important to remember that we are not simply adding the symbol "dx" at the end.
We have seen a practical use of differentials as they offer a good method of making certain approximations. Another use is error
propagation. Suppose a length is measured to be x, although the actual value is x + Δx (where we hope \Delta x\ is small). This
measurement of x may be used to compute some other value; we can think of this as f (x) for some function f . As the true length
is x + Δx , one really should have computed f (x + Δx) . The difference between f (x) and f (x + Δx) is the propagated error.

4.4.3 https://math.libretexts.org/@go/page/4174
How close are f (x) and f (x + Δx) ? This is a difference in "y" values;
$$f(x+\Delta x)-f(x) = \Delta y \approx dy.\]
We can approximate the propagated error using differentials.

Example 4.4.4: Using differentials to approximate propagated error

A steel ball bearing is to be manufactured with a diameter of 2cm. The manufacturing process has a tolerance of ±0.1mm in
the diameter. Given that the density of steel is about 7.85g/cm , estimate the propagated error in the mass of the ball bearing.
3

Solution
The mass of a ball bearing is found using the equation "mass = volume × density." In this situation the mass function is a
product of the radius of the ball bearing, hence it is m = 7.85 π r . The differential of the mass is
4

3
3

$$dm = 31.4\pi r^2 dr.\]


The radius is to be 1cm; the manufacturing tolerance in the radius is mm, or
±0.05 ±0.005 cm. The propagated error is
approximately:
Δm ≈ dm (4.4.13)

2
= 31.4π(1 ) (±0.005) (4.4.14)

= ±0.493g (4.4.15)

Is this error significant? It certainly depends on the application, but we can get an idea by computing the relative error. The
ratio between amount of error to the total mass is
dm 0.493
=± (4.4.16)
4
m 7.85 π
3

0.493
=± (4.4.17)
32.88

= ±0.015, (4.4.18)

or ±1.5.
We leave it to the reader to confirm this, but if the diameter of the ball was supposed to be 10cm, the same manufacturing
tolerance would give a propagated error in mass of ±12.33g, which corresponds to a \textit{percent error} of ±0.188\%.
While the amount of error is much greater ($12.33 > 0.493$), the percent error is much lower.

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). This chapter has put the derivative to yet more uses:
Equation solving (Newton's Method)
Related Rates (furthering our use of the derivative to find instantaneous rates of change)
Optimization (applied extreme values), and
Differentials (useful for various approximations and for something called integration).
In the next chapters, we will consider the "reverse" problem to computing the derivative: given a function f , can we find a function
whose derivative is f ? Be able to do so opens up an incredible world of mathematics and applications.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/
Integrated by Justin Marshall.

This page titled 4.4: Differentials is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et al.
via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

4.4.4 https://math.libretexts.org/@go/page/4174
4.E: Applications of Derivatives (Exercises)
4.1: Newton's Method
Terms and Concepts
1. T/F: Given a function f (x), Newton's Method produces an exact solution to f (x) = 0 .
2. T/F: 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
In Exercises 3-7, the roots of f (x) are known or are easily found. Use 5 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

5. f (x) = x2
+ x − 2, x0 = 0

6. f (x) = x2
− 2, x0 = 1.5

7. f (x) = ln x, x0 = 2

In Exercises 8-11, use Newton's Method to approximate all roots of the given functions accurate to 3 places after the
decimal. If an interval is given, find only the roots that lie in that interval. Use technology to obtain good initial
approximations.
8. f (x) = x3
+ 5x
2
−x −1

9. f (x) = x4
+ 2x
3
− 7x
2
−x +5

10. f (x) = x 17
− 2x
13
− 10 x
8
+ 10 on (−2, 2)

11. f (x) = x 2
cos x + (x − 1) sin x on (−3, 3)

In exercises 12-15, 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.
12. f (x) = x 2
, g(x) = cos x

13. f (x) = x 2
− 1, g(x) = sin x

14. f (x) = e
2
x
, g(x) = cos x

15. f (x) = x, g(x) = tan x on [−6, 6]

16. Why does Newton's Method fail in finding the root of f (x) = x 3
− 3x
2
+ x + 3 when x0 = 1 ?
17. Why does Newton's Method fail in finding the root of f (x) = −17x 4
+ 130 x
3 2
− 301 x + 156x + 156 when x0 = 1 ?

4.2: Related Rates


Terms and Concepts
1. T/F: Implicit differentiation is often used when solving "related rates" type problems.
2. T/F: A study of related rates is part of the standard police officer training.

Problems
3. Water flows onto a flat surface at a rate of 5cm /s forming a circular puddle 10mm deep. How fast is the radius growing when
3

the radius is:


(a) 1 cm?
(b) 10 cm?
(c) 100 cm?

4.E.1 https://math.libretexts.org/@go/page/9972
4. A circular balloon is inflated with air flowing at a rate of 10cm /s. How fast is the radius of the balloon increasing when the
3

radius is:
(a) 1 cm?
(b) 10 cm?
(c) 100 cm?
5. Consider the traffic situation introduced in Example 100. How fast is the “other car” traveling if the officer and the other car are
each 1/2 mile from the intersection, the other car is traveling due west, the officer is traveling north at 50mph, and the radar reading
is −80mph?
6. Consider the traffic situation introduced in Example 100. Calculate how fast the “other car” is traveling in each of the following
situations.
(a) The officer is traveling due north at 50mph and is 1/2 mile from the intersection, while the other car is 1 mile from the
intersection traveling west and the radar reading is −80mph?
(b) The officer is traveling due north at 50mph and is 1 mile from the intersection, while the other car is 1/2 mile from the
intersection traveling west and the radar reading is −80mph?
7. An F-22 aircraft is flying at 500mph with an elevation of 10,000ft on a straight–line path that will take it directly over an anti–
aircraft gun.

How fast must the gun be able to turn to accurately track the aircraft when the plane is:
(a) 1 mile away?
(b) 1/5 mile away?
(c) Directly overhead?
8. An F-22 aircraft is flying at 500mph with an elevation of 100ft on a straight–line path that will take it directly over an anti–
aircraft gun as in Exercise 7 (note the lower elevation here).
How fast must the gun be able to turn to accurately track the aircraft when the plane is:
(a) 1000 feet away?
(b) 100 feet away?
(c) Directly overhead?
9. A 24ft. ladder is leaning against a house while the base is pulled away at a constant rate of 1 ft/s.

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?
(b) 10 feet from the house?
(c) 23 feet from the house?
(d) 24 feet from the house?

4.E.2 https://math.libretexts.org/@go/page/9972
10. A boat is being pulled into a dock at a constant rate of 30ft/min by a winch located 10ft above the deck of the boat.

At what rate is the boat approaching the dock when the boat is:
(a) 50 feet out?
(b) 15 feet out?
(c) 1 foot from the dock?
(d) What happens when the length of rope pulling in the boat is less than 10 feet long?
11. An inverted cylindrical cone, 20ft deep and 10ft across at the top, is being filled with water at a rate of 10ft /min. At what rate
3

is the water rising in the tank when the depth of the water is:
(a) 1 foot?
(b) 10 feet?
(c) 19 feet?
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.

Suppose the man stands directly next to the weight (i.e., a total rope length of 60ft) and begins to walk away at a rate of 2ft/s. How
fast is the weight rising when the man has walked:
(a) 10 feet?
(b) 40 feet?
How far must the man walk to raise the weight all the way to the pulley?
13. Consider the situation described in Exercise 12. Suppose the man starts 40ft from the weight and begins to walk away at a rate
of 2ft/s.
(a) How long is the rope?
(b) How fast is the weight rising after the man has walked 10 feet?
(c) How fast is the weight rising after the man has walked 40 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 100 feet away, a 5' woman tracks the path of the balloon. When
her sightline with the balloon makes a 45 degree angle with the horizontal, she notes the angle is increasing at about 5 /min.

15. A company that produces landscaping materials is dumping sand into a conical pile. The sand is being poured at a rate of 5ft
3
/sec; the physical properties of the sand, in conjunction with gravity, ensure that the cone’s height is roughly 2/3 the length of the
diameter of the circular base.
How fast is the cone rising when it has a height of 30 feet?

4.3: Optimization
Terms and Concepts
1. T/F: An "optimization problem" is essentially an "extreme values" problem in a "story problem" setting.
2. T/F: This section teaches one to find the extreme values of function that have more than one variable.

4.E.3 https://math.libretexts.org/@go/page/9972
Problems
3. Find the maximum product of two numbers (not necessarily integers) that have a sum of 100.
4. Find the minimum sum of two numbers whose product is 500.
5. Find the maximum sum of two numbers whose product is 500.
6. Find the maximum sum of two numbers, each of which is in [0,300] whose product is 500.
7. Find the maximal area of a right triangle with hypotenuse of length 1.
8. A rancher has 1000 feet of fencing in which to construct adjacent, equally sized rectangular pens. What dimensions should these
pens have to maximize the enclosed area?

9. A standard soda can is roughly cylindrical and holds 355cm of liquid. What dimensions should the cylinder be to minimize the
3

material needed to produce the can? Based on your dimensions, determine whether or not the standard can is produced to minimize
the material costs.
10. Find the dimensions of a cylindrical can with a volume of 206in that minimizes the surface area.
3

The “#10 can”is a standard sized can used by the restaurant industry that holds about 206in with a diameter of 6 2/16in and height
3

of 7in. Does it seem these dimensions where chosen with minimization in mind?
11. The United States Postal Service charges more for boxes whose combined length and girth exceeds 108” (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” standard?
12. 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 = kwh
2
for some constant k .

Given a circular log with diameter of 12 inches, what sized beam can be cut from the log with maximum strength?
13. A power line is to be run to an offshore facility in the manner described in Example 104. The offshore facility is 2 miles at sea
and 5 miles along the shoreline from the power plant. It costs $50,000 per mile to lay a power line underground and $80,000 to run
the line underwater.
How much of the power line should be run underground to minimize the overall costs?
14. A power line is to be run to an offshore facility in the manner described in Example 104. The offshore facility is 5 miles at sea
and 2 miles along the shoreline from the power plant. It costs $50,000 per mile to lay a power line underground and $80,000 to run
the line underwater.
How much of the power line should be run underground to minimize the overall costs?
15. A woman throws a stick into a lake for her dog to fetch; the stick is 20 feet down the shore line and 15 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 22ft/s and swims about 1.5ft/s.
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 104 can
be useful.)

4.E.4 https://math.libretexts.org/@go/page/9972
16. A woman throws a stick into a lake for her dog to fetch; the stick is 15 feet down the shore line and 30 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 22ft/s and swims about 1.5ft/s.
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.)
17. What are the dimensions of the rectangle with largest area that can be drawn inside the unit circle?

4.4: Differentials
Terms and Concepts
1. T/F: 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. T/F: the symbols " dx " and " Δx " represent the same concept.
3. T/F: the symbols " dy " and " Δy " represent the same concept.
4. T/F: Differentials are important in the study of integration.
5. How are the differentials and tangent lines related?

Problems
In Exercises 6-17, use differentials to approximate the given value by hand.
6. 2.05 2

7. 5.93 2

8. 5.1 3

9. 6.8 3


− −

10. √16.5
−−
11. √24
−−
12. √63
3


−−
13. √8.5
3

14. sin 3
15. cos 1.5
16. e 0.1

In Exercises 17-29, compute the differential dy .


17. y = x 2
+ 3x − 5

18. y = x 7
−x
5

19. y = 1

4x
2

20. y = (2x + sin x) 2

21. y = x 2
e
3x

22. y = 4

x
4

23. y = 2x

tan x+1

24. y = ln(5x)
25. y = e x
sin x

26. y = cos(sin x)

4.E.5 https://math.libretexts.org/@go/page/9972
27. y = x+1

x+2

28. y = 3 x
ln x

29. y = x ln x − x
30. A set of plastic spheres are to be made with a diameter of 1cm. If the manufacturing process is accurate to 1mm, what is the
propagated error in volume of the spheres?
31. The distance, in feet, a stone drops in t seconds is given by d(t) = 16t . The depth of a hole is to be approximated by dropping
2

a rock and listening for it to hit the bottom. What is the propagated error if the time measurement is accurate to 2/10ths of a second
and the measured time is:
(a) 2 seconds?
(b) 5 seconds?
32. What is the propagated error in the measurement of the cross sectional area of a circular log if the diameter is measured at 15′′,
accurate to 1/4 ′′?
33. A wall is to be painted that is 8′ high and is measured to be 10′ , 7′′ long. Find the propagated error in the measurement of the
wall’s surface area if the measurement is accurate to 1/2 ′′ .
Exercises 34-38 explore some issues related to surveying in which distances are approximated using other measured
distance and measured angles. (Hint: Convert all angle to radians before computing.)
34. The length l of a long wall is to be approximated. The angle θ , as shown in the diagram (not to scale), is measured to be 85.2 , ∘

accurate to 1 . Assume that the triangle formed is a right triangle.


(a) What is the measured length l of the wall?


(b) What is the propagated error?
(c) What is the percent error?
35. Answer the questions to Exercise 34, but with a measured angle of 71.5 , accurate to 1 , measured with a point 100' from the
∘ ∘

wall.
36. The length l of a long wall is to be calculated by measuring the angle θ shown in the diagram (not to scale). Assume the formed
triangle is an isosceles triangle. The measured angle is 143 , accurate to 1 .
∘ ∘

(a) What is the measured length of the wall?


(b) what is the propagated error?
(c) What is the percent error?
37. The length of the walls in Exercise 34-36 are essentially the same. Which setup gives the most accurate result?
38. consider the setup in Exercises 36. This time, assume the angle measurement of 143

is exact but the measured 50

from the
wall is accurate to 6 ". What is the approximate percent error?

4.E: Applications of Derivatives (Exercises) is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by LibreTexts.

4.E.6 https://math.libretexts.org/@go/page/9972
CHAPTER OVERVIEW
5: Integration
5.1: Antiderivatives and Indefinite Integration
5.2: The Definite Integral
5.3: Riemann Sums
5.4: The Fundamental Theorem of Calculus
5.5: Numerical Integration
5.E: Applications of Integration (Exercises)

Contributors
Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 5: Integration is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et al. via
source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

1
5.1: Antiderivatives and Indefinite Integration
We have spent considerable time considering the derivatives of a function and their applications. In the following chapters, we are
going to starting thinking in "the other direction." That is, given a function f (x), we are going to consider functions F (x) such that
F (x) = f (x). There are numerous reasons this will prove to be useful: these functions will help us compute areas, volumes, mass,

force, pressure, work, and much more.


Given a function y = f (x), a differential equation is one that incorporates y , x, and the derivatives of y . For instance, a simple
differential equation is:
$$y' = 2x.\]
Solving a differential equation amounts to finding a function y that satisfies the given equation. Take a moment and consider that
equation; can you find a function y such that y = 2x ?

Can you find another?


And yet another?
Hopefully one was able to come up with at least one solution: y = x . "Finding another" may have seemed impossible until one
2

realizes that a function like y = x + 1 also has a derivative of 2x. Once that discovery is made, finding "yet another" is not
2

difficult; the function y = x + 123, 456, 789 also has a derivative of 2x. The differential equation y = 2x has many solutions.
2 ′

This leads us to some definitions.

Definition 5.1.1: Antiderivatives and Indefinite Integrals

Let a function f (x) be given. An antiderivative of f (x) is a function F (x) such that F ′
.
(x) = f (x)

The set of all antiderivatives of f (x) is the indefinite integral of f , denoted by


$$\int f(x) \ dx.\]

Make a note about our definition: we refer to an antiderivative of f , as opposed to the antiderivative of f , since there is always an
infinite number of them. We often use upper-case letters to denote antiderivatives.
Knowing one antiderivative of f allows us to find infinitely more, simply by adding a constant. Not only does this give us more
antiderivatives, it gives us all of them.

Theorem 5.1.1: Antiderivative Forms

Let F (x) and G(x) be antiderivatives of f (x). Then there exists a constant C such that
$$G(x) = F(x) + C.\]

Given a function f and one of its antiderivatives F , we know all antiderivatives of f have the form F (x) + C for some constant
C . Using Definition 5.1.1, we can say that

$$\int f(x) \ dx = F(x) + C.\]


Let's analyze this indefinite integral notation.

Figure 5.1.1 : Understanding the indefinite integral notation.

5.1.1 https://math.libretexts.org/@go/page/4178
Figure 5.1.1 shows the typical notation of the indefinite integral. The integration symbol, ∫ , is in reality an "elongated S,"
representing "take the sum." We will later see how sums and antiderivatives are related.
The function we want to find an antiderivative of is called the integrand. It contains the differential of the variable we are
integrating with respect to. The ∫ symbol and the differential dx are not "bookends" with a function sandwiched in between;
rather, the symbol ∫ means "find all antiderivatives of what follows," and the function f (x) and dx are multiplied together; the dx
does not "just sit there."
Let's practice using this notation.

Example 5.1.1: Evaluating indefinite integrals

Evaluate ∫ sin x dx.

Solution
We are asked to find all functions F (x) such that F ′
(x) = sin x . Some thought will lead us to one solution: F (x) = − cos x ,
because (− cos x) = sin x .
d

dx

The indefinite integral of sin x is thus − cos x, plus a constant of integration. So:
$$\int \sin x \ dx = -\cos x + C.\]

A commonly asked question is "What happened to the dx ?" The unenlightened response is "Don't worry about it. It just goes
away." A full understanding includes the following.
This process of antidifferentiation is really solving a differential question. The integral
$$\int \sin x\ dx\]
presents us with a differential, dy = sin x dx . It is asking: "What is y ?" We found lots of solutions, all of the form
y = − cos x + C .

Letting dy = sin x dx , rewrite


$$\int \sin x \ dx \quad \text{as}\quad \int dy.\]
This is asking: "What functions have a differential of the form dy ?" The answer 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 antiderivatives 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.

Example 5.1.2: Evaluating indefinite integrals

Evaluate ∫ (3x 2
+ 4x + 5) dx .
Solution
We seek a function F (x) whose derivative is 3x
2
+ 4x + 5 . When taking derivatives, we can consider functions term--by--
term, so we can likely do that here.
What functions have a derivative of 3x ? Some thought will lead us to a cubic, specifically x
2 3
+ C1 , where C is a constant.
1

What functions have a derivative of 4x? Here the x term is raised to the first power, so we likely seek a quadratic. Some
thought should lead us to 2x + C , where C is a constant.
2
2 2

Finally, what functions have a derivative of 5? Functions of the form 5x + C , where C is a constant.
3 3

Our answer appears to be


$$\int (3x^2+4x+5)\ dx = x^3+C_1+2x^2+C_2+5x+C_3.\]

5.1.2 https://math.libretexts.org/@go/page/4178
We do not need three separate constants of integration; combine them as one constant, giving the final answer of
$$\int (3x^2+4x+5)\ dx = x^3+2x^2+5x+C.\]
It is easy to verify our answer; take the derivative of x 3 3
+ 2x + 5x + C and see we indeed get 3x 2
+ 4x + 5 .

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:
$$\frac{d}{dx}\left(\int f(x)\ dx\right) = f(x).\]
Differentiation "undoes" the work done by antidifferentiation.
Theorem 27 gave a list of the derivatives of common functions we had learned at that point. We restate part of that list here to stress
the relationship between derivatives and antiderivatives. This list will also be useful as a glossary of common antiderivatives as we
learn.

Theorem 5.1.2: Derivatives and Antiderivatives

Common Differentiation Rules Common Indefinite Integration Rules

1. d
(cf (x)) = c ⋅ f (x)

dx 1. ∫ c ⋅ f (x) dx = c ⋅ ∫ f (x) dx
2. d

dx
(f (x) ± g(x)) = f (x) ± g (x)
′ ′

2. ∫ (f (x) ± g(x)) dx = ∫ f (x) dx ± ∫ g(x) dx


3. d

dx
(C) = 0
3. ∫ 0 dx = C
4. d

dx
(x) = 1
4. ∫ 1 dx = ∫ dx = x + C
5. d

dx
n
(x ) = n ⋅ x
n−1
5. ∫ x dx = n
x +C
1 n+1
n+1

6. d

dx
( sin x) = cosx 6. ∫ cosx dx = sin x + C
7. d

dx
( cosx) = − sin x 7. ∫ sin x dx = − cosx + C
8. d

dx
( tan x) = sec
2
x 8. ∫ sec x dx = tan x + C
2

9. d
( csc x) = − csc x cot x 9. ∫ csc x cot x dx = − csc x + C
dx

10. d
( sec x) = sec x tan x
10. ∫ sec x tan x dx = sec x + C
dx

11. d
( cot x) = − csc
2
x
11. ∫ csc x dx = − cot x + C
2

dx
12. ∫ e dx = e + C
x x

12. d x
(e ) = e
x

dx
13. ∫ a dx =
x
⋅a +C
1 x

13. d

dx
x
(a ) = ln a ⋅ a
x ln a

14. ∫
1
dx = ln |x| + C
14. d

dx
( ln x) =
1

x
x

We highlight a few important points from Theorem 5.1.2:


Rule #1 states ∫ c ⋅ f (x) dx = c ⋅ ∫ f (x) dx . This is the Constant Multiple Rule: we can temporarily ignore constants when
finding antiderivatives, just as we did when computing derivatives (i.e., (3x ) is just as easy to compute as (x )). An
d

dx
2 d

dx
2

example:

∫ 5 cos x dx = 5 ⋅ ∫ cos x dx = 5 ⋅ (sin x + C ) = 5 sin x + C . (5.1.1)

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 ,".
Rule #2 is the Sum/Difference Rule: we can split integrals apart when the integrand contains terms that are added/subtracted, as
we did in Example 5.1.2. So:

5.1.3 https://math.libretexts.org/@go/page/4178
2 2
∫ (3 x + 4x + 5) dx =∫ 3x dx + ∫ 4x dx + ∫ 5 dx (5.1.2)

2
=3∫ x dx + 4 ∫ x dx + ∫ 5 dx (5.1.3)

1 3
1 2
=3⋅ x +4 ⋅ x + 5x + C (5.1.4)
3 2
3 2
=x + 2x + 5x + C (5.1.5)

In practice we generally do not write out all these steps, but we demonstrate them here for completeness.
Rule #5 is the Power Rule of indefinite integration. There are two important things to keep in mind:
1. Notice the restriction that n ≠ −1 . This is important: ∫ dx ≠ " x + C "; rather, see Rule #14.
1

x
1

0
0

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 antiderivative, do the opposite things in the opposite order: first add one to the power, then second divide by the
power.
Note that Rule #14 incorporates the absolute value of x. The exercises will work the reader through why this is the case; for
now, know the absolute value is important and cannot be ignored.

Initial Value Problems


In Section 2.3 we saw that the derivative of a position function gave a velocity function, and the derivative of a velocity function
describes acceleration. We can now go "the other way:" the antiderivative of an acceleration function gives a velocity function, etc.
While there is just one derivative of a given function, there are infinite antiderivatives. Therefore we cannot ask "What is the
velocity of an object whose acceleration is −32ft/s ?", since there is more than one answer.
2

We can find the answer if we provide more information with the question, as done in the following example. Often the additional
information comes in the form of an initial value, a value of the function that one knows beforehand.

Example 5.1.3: Solving initial value problems

The acceleration due to gravity of a falling object is −32 ft/s . At time t = 3 , a falling object had a velocity of −10 ft/s. Find
2

the equation of the object's velocity.


Solution
We want to know a velocity function, v(t) . We know two things:
1. The acceleration, i.e., v (t) = −32 , and

2. the velocity at a specific time, i.e., v(3) = −10 .


Using the first piece of information, we know that v(t) is an antiderivative of ′
v (t) = −32 . So we begin by finding the
indefinite integral of −32:
$$\int (-32)\ dt = -32t+C=v(t).\]
Now we use the fact that v(3) = −10 to find C :
v(t) = −32t + C (5.1.6)

v(3) = −10 (5.1.7)

−32(3) + C = −10 (5.1.8)

C = 86 (5.1.9)

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 $v(0) = 86$ ft/s. Since the velocity is positive, the object was moving upward.
When did the object begin moving down? Immediately after v(t) = 0 :
$$-32t+86 = 0 \quad \Rightarrow\quad t = \frac{43}{16} \approx 2.69\text{s}.\]

5.1.4 https://math.libretexts.org/@go/page/4178
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.

Example 5.1.4: Solving initial value problems

Find f (t), given that f ′′


(t) = cos t ,f ′
(0) = 3 and f (0) = 5 .
Solution
We start by finding f ′
(t) , which is an antiderivative of f ′′
(t) :
$$\int f''(t)\ dt = \int \cos t\ dt = \sin t + C = f'(t).\]
So f ′
(t) = sin t + C for the correct value of C . We are given that f ′
(0) = 3 , so:
$$f'(0) = 3 \quad \Rightarrow \quad \sin 0+C = 3 \quad \Rightarrow \quad C=3.\]
Using the initial value, we have found f ′
(t) = sin t + 3.

We now find f (t) by integrating again.


$$f(t)=\int f'(t) \ dt = \int (\sin t+3)\ dt = -\cos t + 3t + C.\]
We are given that f (0) = 5 , so
− cos 0 + 3(0) + C =5 (5.1.10)

−1 + C = 5 (5.1.11)

C =6 (5.1.12)

Thus f (t) = − cos t + 3t + 6 .

This section introduced antiderivatives and the indefinite integral. We found they are needed when finding a function given
information about its derivative(s). For instance, we found a position function given a velocity function.
In the next section, we will see how position and velocity are unexpectedly related by the areas of certain regions on a graph of the
velocity function. Then, in Section 5.4, we will see how areas and antiderivatives are closely tied together.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/
Integrated by Justin Marshall.

This page titled 5.1: Antiderivatives and Indefinite Integration is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated
by Gregory Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is
available upon request.

5.1.5 https://math.libretexts.org/@go/page/4178
5.2: The Definite Integral
We start with an easy problem. An object travels in a straight line at a constant velocity of 5 ft/s for 10 seconds. How far away from
its starting point is the object?
We approach this problem with the familiar "Distance = Rate × Time" equation. In this case, Distance = 5ft/s × 10s = 50 feet.
It is interesting to note that this solution of 50 feet can be represented graphically. Consider Figure 5.2.1, where the constant
velocity of 5ft/s is graphed on the axes. Shading the area under the line from t = 0 to t = 10 gives a rectangle with an area of 50
square units; when one considers the units of the axes, we can say this area represents 50 ft.

Figure 5.2.1 : The area under a constant velocity function corresponds to distance traveled.
Now consider a slightly harder situation (and not particularly realistic): an object travels in a straight line with a constant velocity
of 5ft/s for 10 seconds, then instantly reverses course at a rate of 2ft/s for 4 seconds. (Since the object is traveling in the opposite
direction when reversing course, we say the velocity is a constant −2ft/s.) How far away from the starting point is the object --
what is its displacement?
Here we use "Distance = Rate 1 × Time + Rate
1 2 × Time ," which is
2

Distance = 5 ⋅ 10 + (−2) ⋅ 4 = 42 ft. (5.2.1)

Hence the object is 42 feet from its starting location.


We can again depict this situation graphically. In Figure 5.2.3 we have the velocities graphed as straight lines on [0, 10] and
[10, 14], respectively. The displacement of the object is

"Area above the t--axis − Area below the t-axis, (5.2.2)

which is easy to calculate as 50 − 8 = 42 feet.

5.2.1 https://math.libretexts.org/@go/page/4179
Figure 5.2.2 : The total displacement is the area above the t --axis minus the area below the t --axis.
Now consider a more difficult problem.

Example 5.2.1: Finding position using velocity

The velocity of an object moving straight up/down under the acceleration of gravity is given as v(t) = −32t + 48 , where time
t is given in seconds and velocity is in ft/s. When t = 0 , the object had a height of 0 ft.

1. What was the initial velocity of the object?


2. What was the maximum height of the object?
3. What was the height of the object at time t = 2 ?
Solution
It is straightforward to find the initial velocity; at time t = 0 , v(0) = −32 ⋅ 0 + 48 = 48 ft/s.
To answer questions about the height of the object, we need to find the object's position function s(t) . This is an initial value
problem, which we studied in the previous section. We are told the initial height is 0, i.e., s(0) = 0 . We know
s (t) = v(t) = −32t + 48 . To find s , we find the indefinite integral of v(t) :

2
∫ v(t) dt = ∫ (−32t + 48) dt = −16 t + 48t + C = s(t). (5.2.3)

Since s(0) = 0 , we conclude that C =0 and s(t) = −16t 2


+ 48t .
To find the maximum height of the object, we need to find the maximum of s . Recalling our work finding extreme values, we
find the critical points of s by setting its derivative equal to 0 and solving for t :

s (t) = −32t + 48 = 0 ⇒ t = 48/32 = 1.5s. (5.2.4)

(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
2
s(1.5) = −16(1.5 ) + 48(1.5) = 36ft. (5.2.5)

The height at time t = 2 is now straightforward to compute: it is s(2) = 32 ft.


While we have answered all three questions, let's look at them again graphically, using the concepts of area that we explored
earlier.
Figure 5.2.3 shows a graph of v(t) on axes from t = 0 to t = 3 . It is again straightforward to find v(0) . How can we use the
graph to find the maximum height of the object?

Figure 5.2.3 : A graph of v(t) = −32t + 48 ; the shaded areas help determine displacement.
Recall how in our previous work that the displacement of the object (in this case, its height) was found as the area under the
velocity curve, as shaded in the figure. Moreover, the area between the curve and the t --axis that is below the t --axis counted

5.2.2 https://math.libretexts.org/@go/page/4179
as "negative" area. That is, it represents the object coming back toward its starting position. So to find the maximum distance
from the starting point -- the maximum height -- we find the area under the velocity line that is above the t --axis, i.e., from
t = 0 to t = 1.5 . This region is a triangle; its area is

1 1
Area = Base × Height = × 1.5s × 48ft/s = 36ft, (5.2.6)
2 2

which matches our previous calculation of the maximum height.


Finally, we find the total signed area under the velocity function from t = 0 to t = 2 to find the s(2) , the height at t =2 ,
which is a displacement, the distance from the current position to the starting position. That is,

Displacement = Area above the t−axis − Area below t-axis. (5.2.7)

The regions are triangles, and we find


$$\text{Displacement} = \frac12(1.5\text{s})(48\text{ft/s}) - \frac12(.5\text{s})(16\text{ft/s}) = 32\text{ft}.\]
This also matches our previous calculation of the height at t = 2 .
Notice how we answered each question in this example in two ways. Our first method was to manipulate equations using our
understanding of antiderivatives and derivatives. Our second method was geometric: we answered questions looking at a graph
and finding the areas of certain regions of this graph.

The above example does not prove a relationship between area under a velocity function and displacement, but it does imply a
relationship exists. Section 5.4 will fully establish fact that the area under a velocity function is displacement.
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.

Definition 5.2.1: The Definite Integral, Total Signed Area

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 f and above the x-axis on [a, b]). −(area above f and under the x-axis on [a, b]). (5.2.8)

The definite integral of f on [a, b] is the total signed area of f on [a, b], denoted
b

∫ f (x) dx, (5.2.9)


a

There a and b are the bounds of integration.

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."
The previous section introduced the indefinite integral, which related to antiderivatives. We have now defined the definite integral,
which relates to areas under a function. The two are very much related, as we'll see when we learn the Fundamental 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 .
We practice using this notation.

Example 5.2.2: Evaluating definite integrals

Consider the function f given in Figure 5.2.4.

5.2.3 https://math.libretexts.org/@go/page/4179
Figure 5.2.4 : A graph of f (x) in Example 2
Find:
3
1. ∫0
f (x) dx
5
2. ∫3
f (x) dx
5
3. ∫0
f (x) dx
3
4. ∫0
5f (x) dx
1
5. ∫1
f (x) dx

Solution
3 3
1. ∫0
f (x)dx is the area under f on the interval [0, 3]. This region is a triangle, so the area is ∫ f (x)dx = (3)(1) = 1.5 .
0
1

2
5
2. ∫3
f (x)dx represents the area of the triangle found under the x--axis on [3, 5]. The area is
1

2
(2)(1) = 1 ; since it is found
5
under the x--axis, this is "negative area." Therefore ∫ 3
f (x) dx = −1 .
5
3. ∫0
f (x)dxis the total signed area under f on [0, 5]. This is 1.5 + (−1) = 0.5 .
3
4. ∫0
5f (x)dx is the area under 5f on [0, 3]. This is sketched in Figure 5.2.5. Again, the region is a triangle, with height 5
3
times that of the height of the original triangle. Thus the area is ∫ 0
5f (x) dx = 15/2 = 7.5.

Figure 5.2.5 : A graph of 5f in Example 5.2.2. (Yes, it looks just like the graph of f in Figure 5.2.4, just with a different y -
scale.)
1
5. ∫ f (x)dx is the area under f on the "interval" [1, 1]. This describes a line segment, not a region; it has no width.
1

Therefore the area is 0.

This example illustrates some of the properties of the definite integral, given here.

5.2.4 https://math.libretexts.org/@go/page/4179
Theorem 5.2.2: Properties of the Definite Integral

Let f and g be defined on a closed interval I that contains the values a , b and c , and let k be a constant. The following hold:
a
1. ∫a
f (x)dx = 0
b c c
2. ∫a
f (x)dx + ∫
b
f (x)dx = ∫
a
f (x)dx
b a
3. ∫a
f (x)dx = − ∫
b
f (x)dx
b b b
4. ∫a
(f (x) ± g(x))dx = ∫
a
f (x) dx ± ∫
a
g(x)dx
b b
5. ∫a
k ⋅ f (x)dx = k ⋅ ∫
a
f (x)dx

We give a brief justification of Theorem 5.2.2 here.


1. As demonstrated in Example 5.2.2, there is no "area under the curve" when the region has no width; hence this definite integral
is 0.
2. This states that total area is the sum of the areas of subregions. It is easily considered when we let a < b < c . We can break the
interval [a, c] into two subintervals, [a, b] and [b, c]. The total area over [a, c] is the area over [a, b] plus the area over [b, c].
It is important to note that this still holds true even if a < b < c is not true. We discuss this in the next point.
3. This property can be viewed a merely a convention to make other properties work well. (Later we will see how this property has
a justification all its own, not necessarily in support of other properties.) Suppose b < a < c . The discussion from the previous
point clearly justifies
a c c

∫ f (x)dx + ∫ f (x) dx = ∫ f (x)dx. (5.2.10)


b a b

However, we still claim that, as originally stated,


b c c

∫ f (x) dx + ∫ f (x) dx = ∫ f (x) dx. (5.2.11)


a b a

How do Equations 5.2.11 and 5.2.12 relate? Start with Equation 5.2.11:
a c c

∫ f (x)dx + ∫ f (x)dx = ∫ f (x)dx (5.2.12)


b a b
c a c

∫ f (x) dx = − ∫ f (x)dx + ∫ f (x)dx (5.2.13)


a b b

a
Property (3) justifies changing the sign and switching the bounds of integration on the − ∫ f (x)dx term; when this is done, b

Equations 5.2.11 and 5.2.12 are equivalent.


The conclusion is this: by adopting the convention of Property (3), Property (2) holds no matter the order of a , b and c . Again,
in the next section we will see another justification for this property.
4. ,5. Each of these may be non-intuitive. Property (5) states that when one scales a function by, for instance, 7, the area of the
enclosed region also is scaled by a factor of 7. Both Properties (4) and (5) can be proved using geometry. The details are not
complicated but are not discussed here.

Example 5.2.3: Evaluating definite integrals using Theorem 5.2.2.

Consider the graph of a function f (x) shown in Figure 5.2.6.

5.2.5 https://math.libretexts.org/@go/page/4179
Figure 5.2.6 : A graph of a function in Example 5.2.3
Answer the following:
b c
1. Which value is greater: ∫ f (x)dx or ∫
a b
f (x)dx ?
c
2. Is ∫ f (x)dx greater or less than 0?
a
b b
3. Which value is greater: ∫ a
f (x)dx or ∫
c
f (x)dx ?
Solution
b c
1. ∫a
f (x)dx has a positive value (since the area is above the x--axis) whereas ∫ b
f (x)dx has a negative value. Hence
b
∫ f (x)dx is bigger.
a
c
2. ∫ 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
a

than the region above, we conclude that the definite integral has a value less than 0.
b
3. Note how the second integral has the bounds "reversed." Therefore ∫ f (x)dx represents a positive number, greater than
c
b
the area described by the first definite integral. Hence ∫ c
f (x)dx is greater.

The area definition of the definite integral allows us to use geometry compute the definite integral of some simple functions.

Example 5.2.4: Evaluating definite integrals using geometry

Evaluate the following definite integrals:


5 3
− −−− −
2
1. ∫ (2x − 4) dx 2. ∫ √ 9 − x dx. (5.2.14)
−2 −3

Solution

5.2.6 https://math.libretexts.org/@go/page/4179
−−−−−
Figure 5.2.7 : A graph of f (x) = 2x − 4 in (a) and f (x) = √9 − x in (b), from Example 5.2.4
2

1. It is useful to sketch the function in the integrand, as shown in Figure 5.2.7a. We see we need to compute the areas of two
regions, which we have labeled R and R . Both are triangles, so the area computation is straightforward:
1 2

1 1
R1 : (4)(8) = 16 R2 : (3)6 = 9. (5.2.15)
2 2

Region R lies under the x--axis, hence it is counted as negative area (we can think of the triangle's height as being "−8"),
1

so
5

∫ (2x − 4) dx = −16 + 9 = −7. (5.2.16)


−2

2. Recognize that the integrand of this definite integral describes a half circle, as sketched in Figure 5.2.7b, with radius 3.
Thus the area is:
$$\int_{-3}^3 \sqrt{9-x^2}\ dx = \frac12\pi r^2 = \frac 92\pi.\]

Example 5.2.5: Understanding motion given velocity

Consider the graph of a velocity function of an object moving in a straight line, given in Figure 5.2.8, where the numbers in the
given regions gives the area of that region. Assume that the definite integral of a velocity function gives displacement. Find the
maximum speed of the object and its maximum displacement from its starting position.
Solution

Figure 5.2.8 : A graph of a velocity in Example 5.2.5.

5.2.7 https://math.libretexts.org/@go/page/4179
Since the graph gives velocity, finding the maximum speed is simple: it looks to be 15ft/s.
At time t = 0 , the displacement is 0; the object is at its starting position. At time t = a , the object has moved backward 11
feet. Between 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 5.2.9, where a region below y = x is shaded. What is its area? The
2

function y = x is relatively simple, yet the shape it defines has an area that is not simple to find geometrically.
2

Figure 5.2.9 : What is the area below y = x on [0, 3]? The region is not a usual geometric shape.
2

In the next section we will explore how to find the areas of such regions.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/
Integrated by Justin Marshall.

This page titled 5.2: The Definite Integral is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory
Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

5.2.8 https://math.libretexts.org/@go/page/4179
5.3: Riemann Sums
In the previous section we defined the definite integral of a function on [a, b] to be the signed area between the curve and the x--axis. Some areas were simple to
compute; we ended the section with a region whose area was not simple to compute. In this section we develop a technique to find such areas.
A fundamental calculus technique is to first answer a given problem with an approximation, then refine that approximation to make it better, then use limits in the
refining process to find the exact answer. That is exactly what we will do here.
4
Consider the region given in Figure 5.3.1, which is the area under y = 4x − x on [0, 4]. What is the signed area of this region -- i.e., what is ∫
2
0
2
(4x − x )dx ?

Figure 5.3.1 : A graph of f (x) = 4x − x . What is the area of the shaded region?
2

We start by approximating. We can surround the region with a rectangle with height and width of 4 and find the area is approximately 16 square units. This is
obviously an over-approximation; we are including area in the rectangle that is not under the parabola.
We have an approximation of the area, using one rectangle. How can we refine our approximation to make it better? The key to this section is this answer: use more
rectangles.
Let's use 4 rectangles of equal width of 1. This partitions the interval [0, 4] into 4 subintervals, [0, 1], [1, 2], [2, 3] and . On each subinterval we will draw a
[3, 4]

rectangle.
There are three common ways to determine the height of these rectangles: the Left Hand Rule, the Right Hand Rule, and the Midpoint Rule. The Left Hand
Rule says to evaluate the function at the left--hand endpoint of the subinterval and make the rectangle that height. In Figure 5.3.2, the rectangle drawn on the
interval [2, 3] has height determined by the Left Hand Rule; it has a height of f (2). (The rectangle is labeled "LHR.")

4
Figure 5.3.2 : Approximating ∫ 0
2
(4x − x )dx using rectangles. The heights of the rectangles are determined using different rules.
The Right Hand Rule says the opposite: on each subinterval, evaluate the function at the right endpoint and make the rectangle that height. In the figure, the
rectangle drawn on [0, 1] is drawn using f (1) as its height; this rectangle is labeled "RHR.".
The Midpoint Rule says that on each subinterval, evaluate the function at the midpoint and make the rectangle that height. The rectangle drawn on [1, 2] was made
using the Midpoint Rule, with a height of f (1.5). That rectangle is labeled "MPR."
These are the three most common rules for determining the heights of approximating rectangles, but one is not forced to use one of these three methods. The
rectangle on [3, 4] has a height of approximately f (3.53), very close to the Midpoint Rule. It was chosen so that the area of the rectangle is exactly the area of the
region under f on [3, 4]. (Later you'll be able to figure how to do this, too.)
4
The following example will approximate the value of ∫ 0
2
(4x − x )dx using these rules.

Example 5.3.1: Using the Left Hand, Right Hand and Midpoint Rules
4
Approximate the value of ∫ 0
2
(4x − x )dx using the Left Hand Rule, the Right Hand Rule, and the Midpoint Rule, using 4 equally spaced subintervals.
Solution

5.3.1 https://math.libretexts.org/@go/page/4180
{We break the interval [0, 4] into four subintervals as before. In Figure 5.3.3 we see 4 rectangles drawn on f (x) = 4x − x
2
using the Left Hand Rule. (The
areas of the rectangles are given in each figure.)

4
Figure 5.3.3 : Approximating ∫ 0
2
(4x − x )dx using the Left Hand Rule in Example 5.3.1
Note how in the first subinterval, , the rectangle has height
[0, 1] f (0) = 0 . We add up the areas of each rectangle (height× width) for our Left Hand Rule
approximation:

f (0) ⋅ 1 + f (1) ⋅ 1 + f (2) ⋅ 1 + f (3) ⋅ 1 = (5.3.1)

0 +3 +4 +3 = 10. (5.3.2)

Figure 5.3.4 shows 4 rectangles drawn under f using the Right Hand Rule; note how the [3, 4] subinterval has a rectangle of height 0.

4
Figure 5.3.4 : Approximating ∫ 0
2
(4x − x )dx using the Right Hand Rule in Example 5.3.1
In this example, these rectangle seem to be the mirror image of those found in Figure 5.3.3. (This is because of the symmetry of our shaded region.) Our
approximation gives the same answer as before, though calculated a different way:
f (1) ⋅ 1 + f (2) ⋅ 1 + f (3) ⋅ 1 + f (4) ⋅ 1 = (5.3.3)

3 +4 +3 +0 = 10. (5.3.4)

Figure 5.3.5 shows 4 rectangles drawn under f using the Midpoint Rule.

4
Figure 5.3.5 : Approximating ∫ 0
2
(4x − x )dx using the Midpoint Rule in Example 5.3.1
4
This gives an approximation of ∫ 0
2
(4x − x )dx as:
f (0.5) ⋅ 1 + f (1.5) ⋅ 1 + f (2.5) ⋅ 1 + f (3.5) ⋅ 1 = (5.3.5)

1.75 + 3.75 + 3.75 + 1.75 = 11. (5.3.6)

5.3.2 https://math.libretexts.org/@go/page/4180
4
Our three methods provide two approximations of ∫ 0
2
(4x − x )dx : 10 and 11.

Summation Notation
It is hard to tell at this moment which is a better approximation: 10 or 11? We can continue to refine our approximation by using more rectangles. The notation can
become unwieldy, though, as we add up longer and longer lists of numbers. We introduce summation notation to ameliorate this problem.
Suppose we wish to add up a list of numbers a , a , a , \ldots, a . Instead of writing
1 2 3 9

$$a_1+a_2+a_3+a_4+a_5+a_6+a_7+a_8+a_9,\]
we use summation notation and write

Figure 5.3.6 : Understanding summation notation


The upper case sigma represents the term "sum." The index of summation in this example is i; any symbol can be used. By convention, the index takes on only the
integer values between (and including) the lower and upper bounds.
Let's practice using this notation.

Example 5.3.2: Using summation notation

Let the numbers {a } be defined as a


i i = 2i − 1 for integers i, where i ≥ 1 . So a 1 =1 ,a 2 =3 ,a3 =5 , etc. (The output is the positive odd integers). Evaluate
the following summations:
6 7 4

2
1. ∑ ai 2. ∑(3 ai − 4) 3. ∑(ai ) (5.3.7)

i=1 i=3 i=1

Solution

1. 6

∑ ai = a1 + a2 + a3 + a4 + a5 + a6 (5.3.8)

i=1

= 1 + 3 + 5 + 7 + 9 + 11 (5.3.9)

= 36. (5.3.10)

2. Note the starting value is different than 1:


7

∑ ai = (3 a3 − 4) + (3 a4 − 4) + (3 a5 − 4) + (3 a6 − 4) + (3 a7 − 4) (5.3.11)

i=3

= 11 + 17 + 23 + 29 + 35 (5.3.12)

= 115. (5.3.13)

3. 4

2 2 2 2 2
∑(ai ) = (a1 ) + (a2 ) + (a3 ) + (a4 ) (5.3.14)

i=1

2 2 2 2
=1 +3 +5 +7 (5.3.15)

= 84 (5.3.16)

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.

Theorem 5.3.1: Properties of Summations


n
1. ∑ i=1
c =c⋅n , where c is a constant.
n n n
2. ∑ i=m
(ai ± bi ) = ∑
i=m
ai ± ∑
i=m
bi
n n
3. ∑ i=m
c ⋅ ai = c ⋅ ∑
i=m
ai

4. ∑ j

i=m
ai + ∑
n

i=j+1
ai = ∑
n

i=m
ai

n(n+1)
5. ∑ n

i=1
i =
2

n n(n+1)(2n+1)
6. ∑ i=1
2
i =
6
2
n(n+1)
7. ∑ n

i=1
3
i =(
2
)

5.3.3 https://math.libretexts.org/@go/page/4180
Example 5.3.3: Evaluating summations using Theorem5.3.1

Revisit Example 5.3.2 and, using Theorem 5.3.1, evaluate


6 6

∑ ai = ∑(2i − 1). (5.3.17)

i=1 i=1

Solution
6 6 6

∑(2i − 1) = ∑ 2i − ∑(1) (5.3.18)

i=1 i=1 i=1

= (2 ∑ i) − 6 (5.3.19)

i=1

6(6 + 1)
=2 −6 (5.3.20)
2

= 42 − 6 = 36 (5.3.21)

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 5.3.1 is incredibly important when dealing with large sums as we'll soon see.

Riemann Sums
4
Consider again ∫ (4x − x )dx . We will approximate this definite integral using 16 equally spaced subintervals and the Right Hand Rule in Example 5.3.4. Before
0
2

doing so, it will pay to do some careful preparation.

Figure 5.3.7 : Dividing [0, 4] into 16 equally spaced subintervals.


Figure 5.3.7 shows a number line of [0, 4] divided into 16 equally spaced subintervals. We denote 0 as x ; we have marked the values of x , x , x and x . We
1 5 9 13 17

could mark them all, but the figure would get crowded. While it is easy to figure that x = 2.25 , in general, we want a method of determining the value of x
10 i

without consulting the figure. Consider:

So x 10 = x1 + 9(4/16) = 2.25.

If we had partitioned [0, 4] into 100 equally spaced subintervals, each subinterval would have length Δx = 4/100 = 0.04. We could compute x 32 as
$$x_{32} = x_1 + 31(4/100) = 1.24.\]
(That was far faster than creating a sketch first.)
Given any subdivision of [0, 4], the first subinterval is [x 1, ; the second is [x
x2 ] 2, x3 ] ; the i th
subinterval is [x
i, .
xi+1 ]

When using the Left Hand Rule, the height of the i th


rectangle will be f (x ). i

When using the Right Hand Rule, the height of the i th


rectangle will be f (x i+1 ) .
xi +xi+1
When using the Midpoint Rule, the height of the i th
rectangle will be f ( 2
) .
4
Thus approximating ∫ 0
(4x − x )dx
2
with 16 equally spaced subintervals can be expressed as follows, where Δx = 4/16 = 1/4:
Left Hand Rule: ∑ 16

i=1
f (xi )Δx

16
Right Hand Rule: ∑ i=1
f (xi+1 )Δx

xi +xi+1
Midpoint Rule: ∑ 16

i=1
f (
2
) Δx

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 5.3.1

5.3.4 https://math.libretexts.org/@go/page/4180
Example 5.3.4: Approximating definite integrals using sums
4
Approximate ∫ 0
2
(4x − x )dx using the Right Hand Rule and summation formulas with 16 and 1000 equally spaced intervals.
Solution
Using the formula derived before, using 16 equally spaced intervals and the Right Hand Rule, we can approximate the definite integral as
16

∑ f (xi+1 )Δx. (5.3.22)

i=1

We have Δx = 4/16 = 0.25. Since x i = 0 + (i − 1)Δx , we have

xi+1 = 0 + ((i + 1) − 1)Δx (5.3.23)

= iΔx (5.3.24)

Using the summation formulas, consider:


4 16

2
∫ (4x − x )dx ≈ ∑ f (xi+1 )Δx (5.3.25)
0
i=1

16

= ∑ f (iΔx)Δx (5.3.26)

i=1

16
2
= ∑ (4iΔx − (iΔx ) )Δx (5.3.27)

i=1

16

2 2 3
= ∑(4iΔx − i Δx ) (5.3.28)

i=1

16 16

2 3 2
= (4Δx ) ∑ i − Δx ∑i (5.3.29)

i=1 i=1

16 ⋅ 17 16(17)(33)
2 3
= (4Δx ) − Δx (5.3.30)
2 6
2 3
= 4 ⋅ 0.25 ⋅ 136 − 0.25 ⋅ 1496 (5.3.31)

= 10.625 (5.3.32)

We were able to sum up the areas of 16 rectangles with very little computation. In Figure 5.3.8 the function and the 16 rectangles are graphed. While some
rectangles over--approximate the area, other under--approximate the area (by about the same amount). Thus our approximate area of 10.625 is likely a fairly
good approximation.
Notice Equation 5.3.31; by changing the 16's to 1,000's (and appropriately changing the value of Δx), we can use that equation to sum up 1000 rectangles!

4
Figure 5.3.8 : Approximating ∫ 0
2
(4x − x )dx with the Right Hand Rule and 16 evenly spaced subintervals.
We do so here, skipping from the original summand to the equivalent of Equation 5.3.31 to save space. Note that Δx = 4/1000 = 0.004.
4 1000

2
∫ (4x − x )dx ≈ ∑ f (xi+1 )Δx (5.3.33)
0
i=1

1000 1000

2 3 2
= (4Δx ) ∑ i − Δx ∑i (5.3.34)

i=1 i=1

1000 ⋅ 1001 1000(1001)(2001)


2 3
= (4Δx ) − Δx (5.3.35)
2 6
2 3
= 4 ⋅ 0.004 ⋅ 500500 − 0.004 ⋅ 333, 833, 500 (5.3.36)

= 10.666656 (5.3.37)

4
Using many, many rectangles, we have a likely good approximation of ∫ 0
(4x − x )dx
2
. That is,

5.3.5 https://math.libretexts.org/@go/page/4180
$$\int_0^4(4x-x^2)dx \approx 10.666656.\]

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 Δx, 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 rectangle'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 did not have the same size. We refer to the length of the first subinterval as Δx , the length of the 1

second subinterval as Δx , and so on, giving the length of the i subinterval as Δx . Also, one could determine each rectangle's height by evaluating f at
2
th
i

\emph{any} point in the i subinterval. We refer to the point picked in the first subinterval as c , the point picked in the second subinterval as c , and so on, with
th
1 2

c representing the point picked in the i subinterval. Thus the height of the i subinterval would be f (c ), and the area of the i rectangle would be f (c )Δx .
th th th
i i i i

Summations of rectangles with area f (c i )Δxi are named after mathematician Georg Friedrich Bernhard Riemann, as given in the following definition.

Definition 5.3.1: Riemann Sum

Let f be defined on the closed interval [a, b] and let Δx be a partition of [a, b], with
$$a=x_1 < x_2 < \ldots < x_n < x_{n+1}=b.\]
Let Δx denote the length of the i
i
th
subinterval [x i, xi+1 ] and let c denote any value in the i
i
th
subinterval.
The sum
$$\sum_{i=1}^n f(c_i)\Delta x_i\]
is a Riemann sum of f on [a, b].

4
Figure 5.3.9 : An example of a general Riemann sum to approximate ∫ 0
2
(4x − x )dx

4
Figure 5.3.9 shows the approximating rectangles of a Riemann sum of ∫ (4x − x )dx . While the rectangles in this example do not approximate well the shaded
0
2

area, they demonstrate that the subinterval widths may vary and the heights of the rectangles can be determined without following a particular rule.
"Usually" Riemann sums are calculated using one of the three methods we have introduced. The uniformity of construction makes computations easier. Before
working another example, let's summarize some of what we have learned in a convenient way.

Key idea 8: Riemann Sum Concepts


b
Consider ∫ a
f (x)dx ≈ ∑
n

i=1
f (ci )Δxi .

b−a
1. When the n subintervals have equal length, Δx = Δx = . i n

2. The i term of the partition is x = a + (i − 1)Δx . (This makes x


th
i n+1 =b .)
3. The Left Hand Rule summation is: ∑ f (x )Δx. n

i=1 i

4. The Right Hand Rule summation is: ∑ f (x )Δx. n

i=1 i+1

xi +xx+1
5. The Midpoint Rule summation is: ∑ n

i=1
f (
2
) Δx .

Let's do another example.

Example 5.3.5: Approximating definite integrals with sums


3
Approximate ∫ −2
(5x + 2)dx using the Midpoint Rule and 10 equally spaced intervals.
Solution
Following Key Idea 8, we have
3 − (−2)
Δx = = 1/2 and xi = (−2) + (1/2)(i − 1) = i/2 − 5/2. (5.3.38)
10

5.3.6 https://math.libretexts.org/@go/page/4180
xi +xi+1
As we are using the Midpoint Rule, we will also need x i+1 and 2
. Since x i = i/2 − 5/2 ,x i+1 = (i + 1)/2 − 5/2 = i/2 − 2 . This gives

xi + xi+1 (i/2 − 5/2) + (i/2 − 2) i − 9/2


= = = i/2 − 9/4. (5.3.39)
2 2 2

We now construct the Riemann sum and compute its value using summation formulas.
3 10
xi + xi+1
∫ (5x + 2)dx ≈ ∑ f ( ) Δx (5.3.40)
−2
2
i=1

10

= ∑ f (i/2 − 9/4)Δx (5.3.41)

i=1

10

= ∑ (5(i/2 − 9/4) + 2)Δx (5.3.42)

i=1

10
5 37
= Δx ∑ [( )i− ] (5.3.43)
2 4
i=1

10 10
5 37
= Δx ( ∑(i) − ∑ ( )) (5.3.44)
2 4
i=1 i=1

1 5 10(11) 37
= ( ⋅ − 10 ⋅ ) (5.3.45)
2 2 2 4

45
= = 22.5 (5.3.46)
2

Note the graph of f (x) = 5x + 2 in Figure 5.3.10. 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 (c ; when f is
i )Δx

negative, the area is counted as negative.

3
Figure 5.3.10 : Approximating ∫ −2
(5x + 2)dx using the Midpoint Rule and 10 evenly spaced subintervals in Example 5.3.5.

Notice in the previous example that while we used 10 equally spaced intervals, 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 region using n subintervals, where we do not specify a value of n until the very end.

Example 5.3.6: Approximating definite integrals with a formula, using sums


4
Revisit ∫ 0
2
(4x − x )dx yet again. Approximate this definite integral using the Right Hand Rule with n equally spaced subintervals.
Solution
Using Key Idea 8, we know Δx = 4−0

n
= 4/n . We also find x i = 0 + Δx(i − 1) = 4(i − 1)/n . The Right Hand Rule uses x i+1 , which is x
i+1 = 4i/n .
We construct the Right Hand Rule Riemann sum as follows. Be sure to follow each step carefully. If you get stuck, and do not understand how one line
proceeds to the next, you may skip to the result and consider how this result is used. You should come back, though, and work through each step for full
understanding.

5.3.7 https://math.libretexts.org/@go/page/4180
4 n

2
∫ (4x − x )dx ≈ ∑ f (xi+1 )Δx (5.3.47)
0 i=1

n
4i
= ∑f ( ) Δx (5.3.48)
n
i=1

n 2
4i 4i
= ∑ [4 −( ) ] Δx (5.3.49)
n n
i=1

n n
16Δx 16Δx 2
= ∑( ) i −∑( )i (5.3.50)
2
n n
i=1 i=1

n n
16Δx 16Δx 2
=( ) ∑i −( )∑i (5.3.51)
2
n n
i=1 i=1

16Δx n(n + 1) 16Δx n(n + 1)(2n + 1)


=( )⋅ −( ) (recall Δx = 4/n) (5.3.52)
2
n 2 n 6

32(n + 1) 32(n + 1)(2n + 1)


= − (now simplify) (5.3.53)
2
n 3n

32 1
= (1 − ) (5.3.54)
2
3 n

The result is an amazing, easy to use formula. To approximate the definite integral with 10 equally spaced subintervals and the Right Hand Rule, set n = 10

and compute
$$\int_0^4 (4x-x^2)dx \approx \frac{32}{3}\left(1-\frac{1}{10^2}\right) = 10.56.\]
Recall how earlier we approximated the definite integral with 4 subintervals; with n = 4 , the formula gives 10, our answer as before.
It is now easy to approximate the integral with 1,000,000 subintervals! Hand-held calculators will round off the answer a bit prematurely giving an answer of
10.66666667 . (The actual answer is 10.666666666656.)
We now take an important leap. Up to this point, our mathematics has been limited to geometry and algebra (finding areas and manipulating expressions). Now
we apply \textit{calculus}. For any \textit{finite} n , we know that
$$\int_0^4 (4x-x^2)dx \approx \frac{32}{3}\left(1-\frac{1}{n^2}\right).\]
Both common sense and high--level mathematics tell us that as n gets large, the approximation gets better. In fact, if we take the limit as n → ∞ , we get the
4
exact area described by ∫ (4x − x )dx . That is,
0
2

4
32 1
2
∫ (4x − x )dx = lim (1 − ) (5.3.55)
2
0
n→∞ 3 n

32
= (1 − 0) (5.3.56)
3
32 ¯
¯¯
= = 10. 6 (5.3.57)
3

This is a fantastic result. By considering n equally--spaced subintervals, we obtained a formula for an approximation of the definite integral that involved our
variable n . As n grows large -- without bound -- the error shrinks to zero and we obtain the exact area.

This section started with a fundamental calculus technique: make an approximation, refine the approximation to make it better, then use limits in the refining
process to get an exact answer. That is precisely what we just did.
Let's practice this again.

Example 5.3.7: Approximating definite integrals with a formula, using sums


5
Find a formula that approximates ∫ −1
3
x dx using the Right Hand Rule and n equally spaced subintervals, then take the limit as n → ∞ to find the exact area.
Solution
5−(−1)
Following Key Idea 8, we have Δx = n
= 6/n . We have x i = (−1) + (i − 1)Δx ; as the Right Hand Rule uses x
i+1 , we have x
i+1 = (−1) + iΔx .
The Riemann sum corresponding to the Right Hand Rule is (followed by simplifications):

5.3.8 https://math.libretexts.org/@go/page/4180
5 n

3
∫ x dx ≈ ∑ f (xi+1 )Δx (5.3.58)
−1
i=1

= ∑ f (−1 + iΔx)Δx (5.3.59)

i=1

n
3
= ∑(−1 + iΔx ) Δx (5.3.60)

i=1

3 2
= ∑ ((iΔx ) − 3(iΔx ) + 3iΔx − 1)Δx \scriptsize (now distribute Δx) (5.3.61)

i=1

3 4 2 3 2
= ∑ (i Δx − 3 i Δx + 3iΔx − Δx) \scriptsize (now split up summation) (5.3.62)

i=1

n n n n

4 3 3 2 2
= Δx ∑i − 3Δx ∑i + 3Δx ∑ i − ∑ Δx (5.3.63)

i=1 i=1 i=1 i=1

2
n(n + 1) n(n + 1)(2n + 1) n(n + 1)
4 3 2
= Δx ( ) − 3Δx + 3Δx − nΔx (5.3.64)
2 6 2

(use Δx = 6/n) (5.3.65)


2 2
1296 n (n + 1 ) 216 n(n + 1)(2n + 1) 36 n(n + 1)
= ⋅ −3 ⋅ +3 −6 (5.3.66)
4
n 4 n3 6 n2 2

(now do a sizable amount of algebra to simplify) (5.3.67)

378 216
= 156 + + (5.3.68)
n n2

Once again, we have found a compact formula for approximating the definite integral with n equally spaced subintervals and the Right Hand Rule. Using 10
subintervals, we have an approximation of 195.96 (these rectangles are shown in Figure 5.3.11. Using n = 100 gives an approximation of 159.802.

5
Figure 5.3.11 : Approximating ∫ −1
3
x dx using the Right Hand Rule and 10 evenly spaced subintervals.
Now find the exact answer using a limit:
$$\int_{-1}^5 x^3dx = \lim_{n\to\infty} \left(156 + \frac{378}n + \frac{216}{n^2}\right) = 156.\]

Limits of Riemann Sums


We have used limits to evaluate exactly given definite limits. Will this always work? We will show, given not--very--restrictive conditions, that yes, it will always
work.
The previous two examples demonstrated how an expression such as
$$\sum_{i=1}^n f(x_{i+1})\Delta x\]
can be rewritten as an expression explicitly involving n , such as 32/3(1 − 1/n ). 2

Viewed in this manner, we can think of the summation as a function of n . An n value is given (where n is a positive integer), and the sum of areas of n equally
spaced rectangles is returned, using the Left Hand, Right Hand, or Midpoint Rules.
b
Given a definite integral ∫ a
f (x)dx, let:
1. SL (n) =∑
n

i=1
f (xi )Δx , the sum of equally spaced rectangles formed using the Left Hand Rule,
2. SR (n) =∑
n

i=1
f (xi+1 )Δx , the sum of equally spaced rectangles formed using the Right Hand Rule, and
xi +xi+1
3. SM (n) = ∑
n

i=1
f (
2
) Δx , the sum of equally spaced rectangles formed using the Midpoint Rule.

Recall the definition of a limit as n → ∞ : lim n→∞ SL (n) = K if, given any ϵ > 0 , there exists N >0 such that
$$\left|S_L(n)-K\right| < \epsilon \quad \text{when}\quad n\geq N.\]

5.3.9 https://math.libretexts.org/@go/page/4180
b
The following theorem states that we can use any of our three rules to find the exact value of a definite integral ∫ f (x)dx. It also goes two steps further. The
a

theorem states that the height of each rectangle doesn't have to be determined following a specific rule, but could be f (c ), where c is any point in the i
i i
th

subinterval, as discussed before Riemann Sums where defined in Definition 5.3.1.


The theorem goes on to state that the rectangles do not need to be of the same width. Using the notation of Definition 5.3.1, let Δx denote the length of the i
i
th

subinterval in a partition of [a, b]. Now let ||Δx|| represent the length of the largest subinterval in the partition: that is, ||Δx|| is the largest of all the Δx 's. If
i

||Δx|| is small, then [a, b] must be partitioned into many subintervals, since all subintervals must have small lengths. "Taking the limit as ||Δx|| 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
$$\lim_{||\Delta x||\to 0}\sum_{i=1}^nf(c_i)\Delta x_i\]
as "the limit of the sum 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].

Theorem 5.3.2: Definite Integrals and the Limit of Riemann Sums

Let f be continuous on the closed interval [a, b] and let S ,S


L (n) R (n) and S M (n) be defined as before. Then:
1. lim n→∞ SL (n) = limn→∞ SR (n) = limn→∞ SM (n) = limn→∞ ∑
n

i=1
f (ci )Δx ,
b
2. lim n→∞ ∑
n

i=1
f (ci )Δx = ∫
a
f (x)dx$, and .
b
3. lim ∥Δx∥→0

n

i=1
f (ci )Δxi = ∫
a
f (x)dx .

We summarize what we have learned over the past few sections here.
Knowing the "area under the curve" can be useful. One common example is: the area under a velocity curve is displacement.
b
We have defined the definite integral, ∫ f (x)dx, to be the signed area under f on the interval [a, b].
a

While we can approximate a definite integral many ways, we have focused on using rectangles whose heights can be determined using: the Left Hand Rule, the
Right Hand Rule and the Midpoint Rule.
Sums of rectangles of this type are called Riemann sums.
The exact value of the definite integral can be computed using the limit of a Riemann sum. We generally use one of the above methods as it makes the algebra
simpler.
We first learned of derivatives through limits then learned rules that made the process simpler. We know of a way to evaluate a definite integral using limits; in the
next section we will see how the Fundamental Theorem of Calculus makes the process simpler. The key feature of this theorem is its connection between the
indefinite integral and the definite integral.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of VMI and Brian Heinold of Mount
Saint Mary's University. This content is copyrighted by a Creative Commons Attribution - Noncommercial (BY-NC) License. http://www.apexcalculus.com/
Integrated by Justin Marshall.

This page titled 5.3: Riemann Sums is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et al. via source content that was edited
to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

5.3.10 https://math.libretexts.org/@go/page/4180
5.4: The Fundamental Theorem of Calculus
b

Let f (t) be a continuous function defined on [a, b]. The definite integral ∫ f (x) dx is the "area under f " on [a, b]. We can turn
a

this concept into a function by letting the upper (or lower) bound vary.
x

Let F (x) = ∫ f (t) dt . It computes the area under f on [a, x] as illustrated in Figure 5.4.1. We can study this function using our
a
a

knowledge of the definite integral. For instance, F (a) = 0 since ∫ f (t) dt = 0 .


a

Figure 5.4.1 : The area of the shaded region is F (x) = ∫ f (t) dt .


a

We can also apply calculus ideas to F (x); in particular, we can compute its derivative. While this may seem like an innocuous
thing to do, it has far--reaching implications, as demonstrated by the fact that the result is given as an important theorem.

Theorem 5.4.1: The Fundamental Theorem of Calculus, Part 1


x

Let f be continuous on [a, b] and let F (x) = ∫ f (t) dt . Then F is a differentiable function on (a, b), and
a


F (x) = f (x). (5.4.1)

Initially this seems simple, as demonstrated in the following example.

Example 5.4.1: Using the Fundamental Theorem of Calculus, Part 1


x

Let F (x) = ∫ 2
(t + sin t) dt . What is F ′
(x) ?}
−5

Solution
Using the Fundamental Theorem of Calculus, we have F ′
(x) = x
2
+ sin x .

This simple example reveals something incredible: F (x) is an antiderivative of x + sin x ! Therefore, F (x) = x − cos x + C 2 1

3
3

for some value of C . (We can find C , but generally we do not care. We know that F (−5) = 0 , which allows us to compute C . In
this case, C = cos(−5) + .) 125

We have done more than found a complicated way of computing an antiderivative. Consider a function f defined on an open
b x

interval containing a , b and c . Suppose we want to compute ∫ f (t) dt . First, let F (x) = ∫ f (t) dt . Using the properties of the
a c

definite integral found in Theorem 5.2.1, we know

5.4.1 https://math.libretexts.org/@go/page/4181
b c b

∫ f (t) dt = ∫ f (t) dt + ∫ f (t) dt (5.4.2)


a a c

a b

= −∫ f (t) dt + ∫ f (t) dt (5.4.3)


c c

= −F (a) + F (b) (5.4.4)

= F (b) − F (a). (5.4.5)

We now see how indefinite integrals and definite integrals are related: we can evaluate a definite integral using antiderivatives!
This is the second part of the Fundamental Theorem of Calculus.

Theorem 5.4.2: The Fundamental Theorem of Calculus, Part 2

Let f be continuous on [a, b] and let F be any antiderivative of f . Then


b

∫ f (x) dx = F (b) − F (a). (5.4.6)


a

Example 5.4.2: Using the Fundamental Theorem of Calculus, Part 2


4
We spent a great deal of time in the previous section studying ∫ 0
(4x − x ) dx
2
. Using the Fundamental Theorem of Calculus,
evaluate this definite integral.
Solution
We need an antiderivative of f (x) = 4x − x . All antiderivatives of f have the form F (x) = 2x
2 2

1

3
3
x +C ; for simplicity,
choose C = 0 .
The Fundamental Theorem of Calculus states
4
1 64
2 2 3
∫ (4x − x ) dx = F (4) − F (0) = (2(4 ) − 4 ) − (0 − 0) = 32 − = 32/3. (5.4.7)
0
3 3

This is the same answer we obtained using limits in the previous section, just with much less work.

Notation: A special notation is often used in the process of evaluating definite integrals using the Fundamental Theorem of
b

Calculus. Instead of explicitly writing F (b) − F (a) , the notation F (x)




is used. Thus the solution to Example 5.4.2 would be
a

written as:
4 4
1 ∣ 1
2 2 3 2 3
∫ (4x − x ) dx = (2 x − x )∣ = (2(4 ) − 4 ) − (0 − 0) = 32/3. (5.4.8)
0
3 ∣0 3

The Constant C : Any antiderivative F (x) can be chosen when using the Fundamental Theorem of Calculus to evaluate a definite
integral, meaning any value of C can be picked. The constant always cancels out of the expression when evaluating F (b) − F (a) ,
so it does not matter what value is picked. This being the case, we might as well let C = 0 .

Example 5.4.3: Using the Fundamental Theorem of Calculus, Part 2

Evaluate the following definite integrals.


2 π 5 9 5
3 t −
1. ∫ x dx 2. ∫ sin x dx 3. ∫ e dt 4. ∫ √u du 5. ∫ 2 dx (5.4.9)
−2 0 0 4 1

Solution
2 2
1 4∣
1 4 1
1. ∫ x
3
dx = x ∣ =( 2 ) −(
4
(−2 ) ) = 0.
4 ∣ 4 4
−2 −2
π
π

2. ∫ sin x dx = − cos x


= − cos π − ( − cos 0) = 1 + 1 = 2.
0
0

(This is interesting; it says that the area under one "hump" of a sine curve is 2.)

5.4.2 https://math.libretexts.org/@go/page/4181
5
5

3. ∫ t
e dt = e
t∣

∣0
=e
5
−e
0
=e
5
− 1 ≈ 147.41.
0
9 9 9

1 2 3
∣ 2 3 3
2 38
4. ∫ √u du = ∫ u 2
du = u 2
∣ = (9 2
−4 2
) = (27 − 8) = .
3 ∣4 3 3 3
4 4
5
5

5. ∫ 2 dx = 2x


= 2(5) − 2 = 2(5 − 1) = 8.
1
1

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 to 2(4) = 8 .

In general, if c is a constant, then ∫ c dx = c(b − a) .


a

Understanding Motion with the Fundamental Theorem of Calculus


We established, starting with Key Idea 1, that the derivative of a position function is a velocity function, and the derivative of a
velocity function is an acceleration function. Now consider definite integrals of velocity and acceleration functions. Specifically, if
b

v(t) is a velocity function, what does ∫ v(t) dt mean?


a

The Fundamental Theorem of Calculus states that


b

∫ v(t) dt = V (b) − V (a), (5.4.10)


a

where V (t) is any antiderivative of v(t) . Since v(t) is a velocity function, V (t) must be a position function, and V (b) − V (a)

measures a change in position, or displacement.

Example 5.4.4: Finding displacement

A ball is thrown straight up with velocity given by v(t) = −32t + 20 ft/s, where t is measured in seconds. Find, and interpret,
1

∫ v(t) dt. }
0

Solution
Using the Fundamental Theorem of Calculus, we have
1 1

∫ v(t) dt = ∫ (−32t + 20) dt (5.4.11)


0 0

1
2 ∣
= −16 t + 20t (5.4.12)

0

= 4. (5.4.13)

Thus if a ball is thrown straight up into the air with velocity v(t) = −32t + 20 , the height of the ball, 1 second later, will be 4
feet above the initial height. (Note that the ball has traveled much farther. It has gone up to its peak and is falling down, but the
difference between its height at t = 0 and t = 1 is 4 ft.

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. So integrating a speed function 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) = 5 miles/h and t is measured in hours, then 2

∫ a(t) dt = 15 (5.4.14)
0

5.4.3 https://math.libretexts.org/@go/page/4181
means the velocity has increased by 15 m/h from t = 0 to t = 3 .

The Fundamental Theorem of Calculus and the Chain Rule


x

Part 1 of the Fundamental Theorem of Calculus (FTC) states that given F (x) = ∫ f (t) dt ,F ′
(x) = f (x). Using other notation,
a
d

dx
(F (x)) = f (x). While we have just practiced evaluating definite integrals, sometimes finding antiderivatives is impossible and
x

we need to rely on other techniques to approximate the value of a definite integral. Functions written as F (x) = ∫ f (t) dt are
a

useful in such situations.


It may be of further use to compose such a function with another. As an example, we may compose F (x) with g(x) to get
g(x)

F (g(x)) = ∫ f (t) dt. (5.4.15)


a

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). (5.4.16)
dx

An example will help us understand this.

Example 5.4.4: The FTC, Part 1, and the Chain Rule


2
x

Find the derivative of F (x) = ∫ ln t dt .


2

Solution
x

We can view F (x) as being the function G(x) = ∫ ln t dt composed with g(x) = x
2
; that is, F (x) = G(g(x)) . The
2

Fundamental Theorem of Calculus states that G (x) = ln x . The Chain Rule gives us

′ ′ ′
F (x) = G (g(x))g (x) (5.4.17)


= ln(g(x))g (x) (5.4.18)
2
= ln(x )2x (5.4.19)
2
= 2x ln x (5.4.20)

Normally, the steps defining G(x) and g(x) are skipped.

Practice this once more.

Example 5.4.5: The FTC, Part 1, and the Chain Rule


5

Find the derivative of F (x) = ∫ t


3
dt.
cos x

Solution
cos x

Note that F (x) = − ∫ 3


t dt . Viewed this way, the derivative of F is straightforward:
5

′ 3
F (x) = sin x cos x. (5.4.21)

Area Between Curves


Consider continuous functions f (x) and g(x) defined on [a, b], where f (x) ≥ g(x) for all x in , as demonstrated in Figure
[a, b]

5.4.2. What is the area of the shaded region bounded by the two curves over [a, b]?

5.4.4 https://math.libretexts.org/@go/page/4181
Figure 5.4.2 : Finding the area bounded by two functions on an interval; it is found by subtracting 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
b b

∫ f (x) dx − ∫ g(x) dx. (5.4.22)


a a

Properties of the definite integral allow us to simplify this expression to


b

∫ (f (x) − g(x)) dx. (5.4.23)


a

Theorem 5.4.3: Area Between Curves

Let f (x) and g(x) be continuous functions defined on [a, b] where f (x) ≥ g(x) for all x in . The area of the region
[a, b]

bounded by the curves y = f (x), y = g(x) and the lines x = a and x = b is


b

∫ (f (x) − g(x)) dx. (5.4.24)


a

Example 5.4.6: Finding area between curves

Find the area of the region enclosed by y = x 2


+x −5 and y = 3x − 2 .
Solution
It will help to sketch these two functions, as done in Figure 5.4.3.

Figure 5.4.3 : Sketching the region enclosed by y = x 2


+x −5 and y = 3x − 2 in Example 5.4.6

5.4.5 https://math.libretexts.org/@go/page/4181
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 + x − 5 = 3x − 2 and solve for x:
2

2
x +x −5 = 3x − 2 (5.4.25)

2
(x + x − 5) − (3x − 2) =0 (5.4.26)
2
x − 2x − 3 = 0 (5.4.27)

(x − 3)(x + 1) = 0 (5.4.28)

x = −1, 3. (5.4.29)

Following Theorem 5.4.3, the area is


3 3
2 2
∫ (3x − 2 − (x + x − 5)) dx =∫ (−x + 2x + 3) dx (5.4.30)
−1 −1

3
1 3 2

= (− x +x + 3x) ∣ (5.4.31)
3 ∣
−1

1 1
=− (27) + 9 + 9 − ( + 1 − 3) (5.4.32)
3 3

2 ¯
¯¯
= 10 = 10. 6 (5.4.33)
3

The Mean Value Theorem and Average Value

Figure 5.4.4 : A graph of a function f to introduce the Mean Value Theorem.


4

Consider the graph of a function f in Figure 5.4.4 and the area defined by ∫ f (x) dx . Three rectangles are drawn in Figure
1
4

5.4.5; in (a), the height of the rectangle is greater than f on [1, 4], hence the area of this rectangle is is greater than ∫ .
f (x) dx
0

In (b), the height of the rectangle is smaller than f on [1, 4], hence the area of this rectangle is less than ∫ .
f (x) dx
1

5.4.6 https://math.libretexts.org/@go/page/4181
4

Figure 5.4.5 : Differently sized rectangles give upper and lower bounds on ∫ f (x) dx ; the last rectangle matches the area
1

exactly.
4

Finally, in (c) the height of the rectangle is such that the area of the rectangle is exactly that of ∫ f (x) dx . Since rectangles that
0
4

are "too big", as in (a), and rectangles that are "too little," as in (b), give areas greater/lesser than ∫ f (x) dx, it makes sense that
1

there is a rectangle, whose top intersects f (x) somewhere on [1, 4], whose area is exactly that of the definite integral.
We state this idea formally in a theorem.

Theorem 5.4.4: The Mean Value Theorem of Integration

Let f be continuous on [a, b]. There exists a value c in [a, b] such that
b

∫ f (x) dx = f (c)(b − a). (5.4.34)


a

This is an existential statement; c exists, but we do not provide a method of finding it. Theorem 5.4.4 is directly connected to the
Mean Value Theorem of Differentiation, given as Theorem 3.2.1; we leave it to the reader to see how.
We demonstrate the principles involved in this version of the Mean Value Theorem in the following example.

Example 5.4.7: Using the Mean Value Theorem


π

Consider ∫ . Find a value c guaranteed by the Mean Value Theorem.


sin x dx
0

Solution
π

We first need to evaluate ∫ . (This was previously done in Example 5.4.3)


sin x dx
0

π
π

∫ sin x dx = − cos x = 2. (5.4.35)

0
0

Thus we seek a value c in [0, π] such that π sin c = 2 .

π sin c = 2 ⇒ sin c = 2/π ⇒ c = arcsin(2/π) ≈ 0.69. (5.4.36)

Figure 5.4.6 : A graph of y = sin x on [0, π] and the rectangle guaranteed by the Mean Value Theorem.
In Figure 5.4.6 sin x is sketched along with a rectangle with height sin(0.69). The area of the rectangle is the same as the area
under sin x on [0, π].

b b

Let f be a function on [a, b] with c such that f (c)(b − a) = ∫ f (x) dx . Consider ∫ (f (x) − f (c)) dx :
a a

5.4.7 https://math.libretexts.org/@go/page/4181
b b b

∫ (f (x) − f (c)) dx = ∫ f (x) − ∫ f (c) dx (5.4.37)


a a a

= f (c)(b − a) − f (c)(b − a) (5.4.38)

= 0. (5.4.39)

When f (x) is shifted by −f (c), the amount of area under f above the x-axis on [a, b] is the same as the amount of area below the
x-axis above f ; see Figure 5.4.7 for an illustration of this. In this sense, we can say that f (c) is the average value of f on [a, b].

Figure 5.4.7 : 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 the Mean Value Theorem can be rewritten as
b
1
f (c) = ∫ f (x) dx, (5.4.40)
b −a a

for some value of c in [a, b]. Next, partition the interval [a, b] into n equally spaced subintervals, a = x 1 < x2 < … < xn+1 = b

and choose any c in [x , x ]. The average of the numbers f (c ), f (c ), …, f (c ) is:


i i i+1 1 2 n

n
1 1
(f (c1 ) + f (c2 ) + … + f (cn )) = ∑ f (ci ). (5.4.41)
n n
i=1

(b−a)
Multiply this last expression by 1 in the form of (b−a)
:
n n
1 1
∑ f (ci ) = ∑ f (ci ) (5.4.42)
n n
i=1 i=1

n
1 (b − a)
= ∑ f (ci ) (5.4.43)
n (b − a)
i=1

n
1
= ∑ f (ci ) Δx (where Δx = (b − a)/n) (5.4.44)
b −a
i=1

Now take the limit as n → ∞ :


n b
1 1
lim ∑ f (ci ) Δx = ∫ f (x) dx = f (c). (5.4.45)
n→∞ b −a b −a a
i=1

This tells us this: when we evaluate f at n (somewhat) equally spaced points in [a, b], the average value of these samples is f (c) as
n → ∞.

This leads us to a definition.

Definition 5.4.1: The Average Value of f on [a, b]

Let f be continuous on [a, b]. The average value of f on [a, b] is f (c), where c is a value in [a, b] guaranteed by the Mean
Value Theorem. I.e.,

5.4.8 https://math.libretexts.org/@go/page/4181
b
1
Average Value of f on [a, b] = ∫ f (x) dx. (5.4.46)
b −a a

An application of this definition is given in the following example.

Example 5.4.8: 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:
3 3 3
1 1 1 1 ∣
2 2 3 2
∫ (t − 1 ) dt = ∫ (t − 2t + 1) dt = ( t −t + t) ∣ = 1 ft/s. (5.4.47)
3 −0 0 3 0 3 3 ∣
0

We can understand the above example through a simpler situation. Suppose you drove 100 miles in 2 hours. What was your
average speed? The answer is simple: displacement/time = 100 miles/2 hours = 50mph.
What was the displacement of the object in Example 5.4.8? We calculate this by integrating its velocity function:
3


2
(t − 1 ) dt = 3 ft. Its final position was 3 feet from its initial position after 3 seconds: its average velocity was 1 ft/s.
0

This section has laid the groundwork for a lot of great mathematics to follow. The most important lesson is this: definite integrals
can be evaluated using antiderivatives. Since the previous section established that definite integrals are the limit of Riemann sums,
we can later create Riemann sums to approximate values other than "area under the curve," convert the sums to definite integrals,
then evaluate these using the Fundamental Theorem of Calculus. This will allow us to compute the work done by a variable force,
the volume of certain solids, the arc length of curves, and more.
The downside is this: generally speaking, computing antiderivatives is much more difficult than computing derivatives. The next
chapter is devoted to techniques of finding antiderivatives so that a wide variety of definite integrals can be evaluated. Before that,
the next section explores techniques of approximating the value of definite integrals beyond using the Left Hand, Right Hand and
Midpoint Rules.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/
Integrated by Justin Marshall.

This page titled 5.4: The Fundamental Theorem of Calculus is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated
by Gregory Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is
available upon request.

5.4.9 https://math.libretexts.org/@go/page/4181
5.5: Numerical Integration
The Fundamental Theorem of Calculus gives a concrete technique for finding the exact value of a definite integral. That technique
is based on computing antiderivatives. Despite the power of this theorem, there are still situations where we must approximate the
value of the definite integral instead of finding its exact value. The first situation we explore is where we cannot compute the
antiderivative of the integrand. The second case is when we actually do not know the integrand, but only its value when evaluated
at certain points.
An elementary function is any function that is a combination of polynomials, n roots, rational, exponential, logarithmic and
th

trigonometric functions. We can compute the derivative of any elementary function, but there are many elementary functions of
which we cannot compute an antiderivative. For example, the following functions do not have antiderivatives that we can express
with elementary functions:
$$e^{-x^2}, \quad \sin(x^3)\quad \text{and} \quad \frac{\sin x}{x}.\]
The simplest way to refer to the antiderivatives of e is to simply write ∫ e .
2 2
−x −x
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:
$$ \int_0^1 e^{-x^2} \ dx, \quad \int_{-\frac{\pi}{4}}^{\frac{\pi}{2}} \sin(x^3) \ dx, \quad \text{and} \quad \int_{0.5}^{4\pi}
\frac{\sin(x)}{x} \ dx,\]
as pictured in Figure 5.5.1.

Figure 5.5.1 : Graphically representing three definite integrals that cannot be evaluated using antiderivatives.

The Left and Right Hand Rule Methods


In Section 5.3 we addressed the problem of evaluating definite integrals by approximating the area under the curve using
rectangles. We revisit those ideas here before introducing other methods of approximating definite integrals.
b
We start with a review of notation. Let f be a continuous function on the interval [a, b]. We wish to approximate ∫ f (x) dx. We a
b−a
partition [a, b] into n equally spaced subintervals, each of length dx = . The endpoints of these subintervals are labeled as
n

$$x_1=a,\ x_2 = a+dx,\ x_3 = a+ 2dx,\ \ldots,\ x_i = a+(i-1)\ dx,\ \ldots,\ x_{n+1} = b.\]
n
Key Idea 8 states that to use the Left Hand Rule we use the summation ∑ i=1
f (xi ) dx and to use the Right Hand Rule we use
n
∑ f (x i+1) dx . We review the use of these rules in the context of examples.
i=1

Example 5.5.1: Approximating definite integrals with rectangles


1 2

Approximate ∫ 0
e
−x
dx using the Left and Right Hand Rules with 5 equally spaced subintervals.
Solution
We begin by partitioning the interval [0, 1] into 5 equally spaced intervals. We have dx =
1−0

5
= 1/5 = 0.2 , so $$x_1 = 0,\
x_2 = 0.2,\ x_3 = 0.4,\ x_4 = 0.6,\ x_5 = 0.8,\ \text{and}\ x_6 = 1.\]

5.5.1 https://math.libretexts.org/@go/page/4182
Using the Left Hand Rule, we have:
n

∑ f (xi ) dx = (f (x1 ) + f (x2 ) + f (x3 ) + f (x4 ) + f (x5 )) dx (5.5.1)

i=1

= (f (0) + f (0.2) + f (0.4) + f (0.6) + f (0.8)) dx (5.5.2)

≈ (1 + 0.961 + 0.852 + 0.698 + 0.527)(0.2) (5.5.3)

≈ 0.808. (5.5.4)

Using the Right Hand Rule, we have:


n

∑ f (xi+1 ) dx = (f (x2 ) + f (x3 ) + f (x4 ) + f (x5 ) + f (x6 )) dx (5.5.5)

i=1

= (f (0.2) + f (0.4) + f (0.6) + f (0.8) + f (1)) dx (5.5.6)

≈ (0.961 + 0.852 + 0.698 + 0.527 + 0.368)(0.2) (5.5.7)

≈ 0.681. (5.5.8)

1
Figure 5.5.2 : Approximating ∫ in Example 5.5.1
2
−x
e dx
0

Figure 5.5.2 shows the rectangles used in each method to approximate the definite integral. These graphs show that in this
particular case, the Left Hand Rule is an over approximation and the Right Hand Rule is an under approximation. To get a
better approximation, we could use more rectangles, as we did in Section 3.1. We could also average the Left and Right Hand
Rule results together, giving
$$ \frac{0.808 + 0.681}{2} = 0.7445.\]
The actual answer, accurate to 4 places after the decimal, is 0.7468, showing our average is a good approximation.

Example 5.5.2: Approximating definite integrals with rectangles


π

Approximate ∫ −
2
π
3
sin(x ) dx using the Left and Right Hand Rules with 10 equally spaced subintervals.
4

Solution
We begin by finding \Delta x:
$$\frac{b-a}{n} = \frac{\pi/2 - (-\pi/4)}{10} = \frac{3\pi}{40}\approx 0.236.\]
It is useful to write out the endpoints of the subintervals in a table; in Table 5.5.1, we give the exact values of the endpoints,
their decimal approximations, and decimal approximations of sin(x ) evaluated at these points.
3

Table 5.5.1 : Table of values used to approximate ∫ 2


π
3
sin( x ) dx in Example 5.5.1 .
4

xi Exact Approx. 3
sin(x )
i

x1 −π/4 −0.785 −0.466

x2 −7π/40 −0.550 −0.165

x3 −π/10 −0.314 −0.031

5.5.2 https://math.libretexts.org/@go/page/4182
xi Exact Approx. sin(x )
3
i

x4 −π/40 −0.0785 0

x5 π/20 0.157 0.004

x6 π/8 0.393 0.061

x7 π/5 0.628 0.246

x8 11π/40 0.864 0.601

x9 7π/20 1.10 0.971

x10 17π/40 1.34 0.690

x11 π/2 1.57 −0.670

Once this table is created, it is straightforward to approximate the definite integral using the Left and Right Hand Rules. (Note:
the table itself is easy to create, especially with a standard spreadsheet program on a computer. The last two columns are all
that are needed.) The Left Hand Rule sums the first 10 values of sin(x ) and multiplies the sum by dx; the Right Hand Rule
3
i

sums the last 10 values of sin(x ) and multiplies by dx. Therefore we have:
3
i

Left Hand Rule: ∫ −


2
π
3
sin(x ) dx ≈ (1.91)(0.236) = 0.451.
4

Right Hand Rule: ∫ −


2
π
3
sin(x ) dx ≈ (1.71)(0.236) = 0.404.
4

Average of the Left and Right Hand Rules: 0.4275.

Figure 5.5.3 : ∫ 2


π
3
sin(x ) dx in Example 5.5.2
4

The actual answer, accurate to 3 places after the decimal, is 0.460. Our approximations were once again fairly good. The
rectangles used in each approximation are shown in Figure 5.5.3. It is clear from the graphs that using more rectangles (and
hence, narrower rectangles) should result in a more accurate approximation.

The Trapezoidal Rule


1
In Example 5.5.1 we approximated the value of ∫ e dx with 5 rectangles of equal width. Figure 5.5.2 shows the rectangles
2
−x

used in the Left and Right Hand Rules. These graphs clearly show that rectangles do not match the shape of the graph all that well,
and that accurate approximations will only come by using lots of rectangles.
Instead of using rectangles to approximate the area, we can instead use trapezoids. In Figure 5.5.4, we show the region under
on [0, 1] approximated with 5 trapezoids of equal width; the top "corners" of each trapezoid lies on the graph of f (x).
2
−x
f (x) = e

It is clear from this figure that these trapezoids more accurately approximate the area under f and hence should give a better
1
approximation of ∫ e dx . (In fact, these trapezoids seem to give a great approximation of the area!)
2
−x
0

5.5.3 https://math.libretexts.org/@go/page/4182
1
Figure 5.5.4 : Approximating ∫ using 5 trapezoids of equal widths.
2
−x
e dx
0

1
The formula for the area of a trapezoid is given in Figure 5.5.5. We approximate ∫ with these trapezoids in the following
2
−x
e dx
0

example.

Example 5.5.3: Approximating definite integrals using trapezoids


1
Use 5 trapezoids of equal width to approximate ∫ .
2
−x
e dx
0

Solution
To compute the areas of the 5 trapezoids in Figure 5.5.6, it will again be useful to create a table of values as shown in Table
5.5.2.

Figure 5.5.5 : The area of a trapezoid


Table 5.5.2 : A table of values of e .
2
−x

2
−x
xi e i

0 1

0.2 0.961

0.4 0.852

0.6 0.698

0.8 0.527

1 0.368

The leftmost trapezoid has legs of length 1 and 0.961 and a height of 0.2. Thus, by our formula, the area of the leftmost
trapezoid is:
$$ \frac{1+0.961}{2}(0.2) = 0.1961.\]
Moving right, the next trapezoid has legs of length 0.961 and 0.852 and a height of 0.2. Thus its area is:
$\frac{0.961+0.852}2(0.2) = 0.1813.\]
The sum of the areas of all 5 trapezoids is:
1 + 0.961 0.961 + 0.852 0.852 + 0.698
(0.2) + (0.2) + (0.2) + (5.5.9)
2 2 2
0.698 + 0.527 0.527 + 0.368
(0.2) + (0.2) = 0.7445. (5.5.10)
2 2

5.5.4 https://math.libretexts.org/@go/page/4182
1
We approximate ∫
2
−x
e dx ≈ 0.7445.
0

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
dx = 0.2 . We can factor these coefficients out, leaving a more concise summation as:

$$\frac12(0.2)\Big[(1+0.961) + (0.961+0.852) + (0.852+0.698) + ( 0.698+ 0.527) +(0.527 + 0.368)\Big].\]


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
$$\frac{0.2}{2}\Big[1 + 2(0.961+0.852+0.698+0.527) + 0.368\Big].\]
b
This is the heart of the Trapezoidal Rule, wherein a definite integral ∫ f (x) dx is approximated by using trapezoids of equal
a

widths to approximate the corresponding area under f . Using n equally spaced subintervals with endpoints x , x , …, x , we 1 2 n+1

again have Δx = b−a

n
. Thus:
b n
f (xi ) + f (xi+1 )
∫ f (x) dx ≈ ∑ dx (5.5.11)
a
2
i=1

n
dx
= ∑ (f (xi ) + f (xi+1 )) (5.5.12)
2
i=1

n
dx
= [f (x1 ) + 2 ∑ f (xi ) + f (xn+1 )]. (5.5.13)
2
i=2

Example 5.5.4: Using the Trapezoidal Rule


π

Revisit Example 5.5.4 and approximate ∫ −


2
π
3
sin(x ) dx using the Trapezoidal Rule and 10 equally spaced subintervals.
4

Solution
We refer back to Table 5.5.1 for the table of values of sin(x ). Recall that dx = 3π/40 ≈ 0.236. Thus we have:
3

2
3
0.236
∫ sin(x ) dx ≈ [ − 0.466 + 2( − 0.165 + (−0.031) + … + 0.69) + (−0.67)]

π
2
4

= 0.4275.

Notice how "quickly" the Trapezoidal Rule can be implemented once the table of values is created. This is true for all the methods
explored in this section; the real work is creating a table of x and f (x ) values. Once this is completed, approximating the definite
i i

integral is not difficult. Again, using technology is wise. Spreadsheets can make quick work of these computations and make using
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 learning about definite
integrals, but if a real approximation is needed, one is generally better off using the Trapezoidal Rule instead of either the Left or
Right Hand Rule.
How can we improve on the Trapezoidal Rule, apart from using more and more trapezoids? The answer is clear once we look back
and consider what we have really done so far. The Left Hand Rule is not really about using rectangles to approximate area. Instead,
it approximates a function f with constant functions on small subintervals and then computes the definite integral of these constant
functions. The Trapezoidal Rule is really approximating a function f with a linear function on a small subinterval, then computes
the definite integral of this linear function. In both 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 function 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 Simpson's Rule, named after Thomas Simpson (1710-1761), even though others had
used this rule as much as 100 years prior.

5.5.5 https://math.libretexts.org/@go/page/4182
Simpson's Rule
Given one point, we can create a constant function that goes through that point. Given two points, we can create a linear function
that goes through those points. Given three points, we can create a quadratic function that goes through those three points (given
that no two have the same x--value).
Consider three points (x , y ), (x , y ) and (x , y ) whose x--values are equally spaced and x
1 1 2 2 3 3 1 < x2 < x3 . Let f be the quadratic
function that goes through these three points. It is not hard to show that
x3
x3 − x1
∫ f (x) dx = (y1 + 4 y2 + y3 ). (5.5.14)
x1
6

Consider Figure 5.5.6. A function f goes through the 3 points shown and the parabola g that also goes through those points is
graphed with a dashed line. Using our equation from above, we know exactly that
3
3 −1
∫ g(x) dx = (3 + 4(1) + 2) = 3. (5.5.15)
1
6

Since g is a good approximation for f on [1, 3], we can state that $$\int_1^3 f(x)\ dx \approx 3.\]

Figure 5.5.6 : A graph of a function f and a parabola that approximates it well on [1, 3].
Notice how the interval [1, 3] was split into two subintervals as we needed 3 points. Because of this, whenever we use Simpson's
Rule, we need to break the interval into an even number of subintervals.
b
In general, to approximate ∫ f (x) dx using Simpson's Rule, subdivide [a, b] into n subintervals, where n is even and each
a

subinterval has width dx = (b − a)/n . We approximate f with n/2 parabolic curves, using Equation 5.5.22 to compute the area
under these parabolas. Adding up these areas gives the formula:
$$\int_a^b f(x) \ dx \approx \frac{\ dx}3\Big[f(x_1)+4f(x_2)+2f(x_3)+4f(x_4)+\ldots+2f(x_{n-1})+4f(x_n)+f(x_{n+1})\Big].\]
Note how the coefficients of the terms in the summation have the pattern 1, 4, 2, 4, 2, 4, …, 2, 4, 1.
Let's demonstrate Simpson's Rule with a concrete example.

Example 5.5.5: Using Simpson's Rule


1
Approximate ∫ using Simpson's Rule and 4 equally spaced subintervals.
2
−x
e dx
0

Solution
We begin by making a table of values as we have in the past, as shown in Table 5.5.3.
Table 5.5.3 : A table of values to approximate ∫ , along with a graph of the function.
1 −x
2

e dx
0

−x2
xi e i

0 1

0.25 0.939

0.5 0.779

5.5.6 https://math.libretexts.org/@go/page/4182
2
−x
xi e i

0.75 0.570

1 0.368

Simpson's Rule states that


$$\int_0^1e^{-x^2}\ dx \approx \frac{0.25}{3}\Big[1+4(0.939)+2(0.779)+4(0.570) + 0.368\Big] = 0.7468\overline{3}.\]
Recall in Example 5.5.1 we stated that the correct answer, accurate to 4 places after the decimal, was 0.7468. Our
approximation with Simpson's Rule, with 4 subintervals, is better than our approximation with the Trapezoidal Rule using 5!

1
Figure 5.5.7 : Using Simpson's Rule with n = 4 to approximate ∫ .
2
−x
e dx
0

Figure 5.5.7 shows f (x) = e along with its approximating parabolas, demonstrating how good our approximation is. The
2
−x

approximating curves are nearly indistinguishable from the actual function.

Example 5.5.6: Using Simpson's Rule


π

Approximate ∫ −
2
π
3
sin(x ) dx using Simpson's Rule and 10 equally spaced intervals.
4

Solution
Table 5.5.4 shows the table of values that we used in the past for this problem, shown here again for convenience. Again,
dx = (π/2 + π/4)/10 ≈ 0.236 .
π

Table 5.5.4 : Table of values used to approximate ∫



2
π
3
sin( x ) dx in Example 5.5.6 .
4

3
xi sin(x )
i

−0.785 −0.466

−0.550 −0.165

−0.314 −0.031

−0.0785 0

0.157 0.004

0.393 0.061

0.628 0.246

0.864 0.601

1.10 0.971

1.34 0.690

1.57 −0.670

Simpson's Rule states that

5.5.7 https://math.libretexts.org/@go/page/4182
π

2
3
0.236
∫ sin(x ) dx ≈ [(−0.466) + 4(−0.165) + 2(−0.031) + … (5.5.16)

π
3
4

… + 2(0.971) + 4(0.69) + (−0.67)] (5.5.17)

= 0.4701 (5.5.18)

Figure 5.5.8 : Approximating ∫ 2


π
3
sin(x ) dx in Example 5.5.6 with Simpson's Rule and 10 equally spaced intervals.
4

Recall that the actual value, accurate to 3 decimal places, is 0.460. Our approximation is within one 1/100 th
of the correct
value. The graph in Figure 5.5.8 shows how closely the parabolas match the shape of the graph.

Summary and Error Analysis


We summarize the key concepts of this section thus far in the following Key Idea.

Key Idea 9: Numerical Integration

Let f be a continuous function on [a, b], let n be a positive integer, and let Δx = b−a

n
.
Set x1 =a ,x 2 = a + dx , …, x i = a + (i − 1) dx ,x n+1 =b .
b
Consider ∫ a
f (x) dx .
b
Left Hand Rule: ∫ a
f (x) dx ≈ dx[f (x1 ) + f (x2 ) + … + f (xn )] .
b
Right Hand Rule: ∫ a
f (x) dx ≈ dx[f (x2 ) + f (x3 ) + … + f (xn+1 )] .
b
Trapezoidal Rule: ∫ a
f (x) dx ≈
dx

2
[f (x1 ) + 2f (x2 ) + 2f (x3 ) + … + 2f (xn ) + f (xn+1 )] .
b
Simpson's Rule: ∫ a
f (x) dx ≈
dx

3
[f (x1 ) + 4f (x2 ) + 2f (x3 ) + … + 4f (xn ) + f (xn+1 )] (n even)}.

In our examples, we approximated the value of a definite integral using a given method then compared it to the "right" answer. This
should have raised several questions in the reader's mind, such as:
1. How was the "right" answer computed?
2. If the right answer can be found, what is the point of approximating?
3. If there is value to approximating, how are we supposed to know if the approximation is any good?
These are good questions, and their answers are educational. In the examples, the right answer was never computed. Rather, an
approximation accurate to a certain number of places after the decimal was given. In Example 5.5.1, we do not know the exact
answer, but we know it starts with 0.7468. These more accurate approximations were computed using numerical integration but
with more precision (i.e., more subintervals and the help of a computer).
Since the exact answer cannot be found, approximation still has its place. How are we to tell if the approximation is any good?
"Trial and error" provides one way. Using technology, make an approximation with, say, 10, 100, and 200 subintervals. This likely
will not take much time at all, and a trend should emerge. If a trend does not emerge, try using yet more subintervals. Keep in mind
that trial and error is never foolproof; you might stumble upon a problem in which a trend will not emerge.

5.5.8 https://math.libretexts.org/@go/page/4182
A second method is to use Error Analysis. While the details are beyond the scope of this text, there are some formulas that give
bounds for how good your approximation will be. For instance, the formula might state that the approximation is within 0.1 of the
correct answer. If the approximation is 1.58, then one knows that the correct answer is between 1.48 and 1.68. By using lots of
subintervals, one can get an approximation as accurate as one likes. Theorem 5.5.1 states what these bounds are.

Theorem 5.5.1: Error Bounds in the Trapezoidal and Simpson's Rules


b
1. Let E be the error in approximating ∫ f (x) dx using the Trapezoidal Rule.
T a

If f has a continuous 2 derivative on [a, b] and M is any upper bound of ∣∣f " (x)∣∣ on [a, b], then
nd

$$ E_T \leq \frac{(b-a)^3}{12n^2}M.\]


b
2. Let E be the error in approximating ∫ f (x) dx using Simpson's Rule.
S a

If f has a continuous 4 derivative on [a, b] and M is any upper bound of ∣∣f


th (4)

∣ on [a, b], then
$$E_S \leq \frac{(b-a)^5}{180n^4}M.\]

There are some key things to note about this theorem.


1. The larger the interval, the larger the error. This should make sense intuitively.
2. The error shrinks as more subintervals are used (i.e., as n gets larger).
3. The error in Simpson's Rule has a term relating to the 4 derivative of f . Consider a cubic polynomial: it's 4 derivative is 0.
th th

Therefore, the error in approximating the definite integral of a cubic polynomial with Simpson's Rule is 0 -- Simpson's Rule
computes the exact answer!
We revisit Examples 5.5.3 and 5.5.5 and compute the error bounds using Theorem 5.5.1 in the following example.

Example 5.5.7: Computing error bounds


1
Find the error bounds when approximating using the Trapezoidal Rule and 5 subintervals, and using Simpson's
2
−x
∫ e dx
0

Rule with 4 subintervals.


Solution
Trapezoidal Rule with n = 5 :
We start by computing the 2 derivative of f (x) = e :
2
nd −x

2
−x 2
f " (x) = e (4 x − 2). (5.5.19)

Figure 5.5.8shows a graph of f " (x) on [0, 1]. It is clear that the largest value of f " , in absolute value, is 2. Thus we let
M =2 and apply the error formula from Theorem 5.5.1.

Figure 5.5.9 : Graphing f " (x) in Example 5.5.7 to help establish error bounds.
$$E_T = \frac{(1-0)^3}{12\cdot 5^2}\cdot 2 = 0.00\overline{6}.\]
Our error estimation formula states that our approximation of 0.7445 found in Example 5.5.3 is within 0.0067 of the correct
answer, hence we know that

5.5.9 https://math.libretexts.org/@go/page/4182
$$0.7445-0.0067 = .7378 \leq \int_0^1e^{-x^2}\ dx \leq 0.7512 = 0.7445 + 0.0067.\]
We had earlier computed the exact answer, correct to 4 decimal places, to be 0.7468, affirming the validity of Theorem 5.5.1.
Simpson's Rule with n = 4 :
2

We start by computing the 4 th


derivative of f (x) = e −x
:
$$f\,^{(4)}(x) = e^{-x^2}(16x^4-48x^2+12).\]
Figure 5.5.9 shows a graph of f (x) on [0, 1]. It is clear that the largest value of f
(4) (4)
, in absolute value, is 12. Thus we let
M = 12 and apply the error formula from Theorem 5.5.1.

$$E_s = \frac{(1-0)^5}{180\cdot 4^4}\cdot 12 = 0.00026.\]

Figure 5.5.10 : Graphing f (4)


(x) in Example 5.5.7 to help establish error bounds.
¯
¯¯
Our error estimation formula states that our approximation of 0.74683 found in Example 5.5.5 is within 0.00026 of the correct
answer, hence we know that
$$0.74683-0.00026 = .74657 \leq \int_0^1e^{-x^2}\ dx \leq 0.74709 = 0.74683 + 0.00026.\]
Once again we affirm the validity of Theorem5.5.1.

At the beginning of this section we mentioned two main situations where numerical integration was desirable. We have considered
the case where an antiderivative of the integrand cannot be computed. We now investigate the situation where the integrand is not
known. This is, in fact, the most widely used application of Numerical Integration methods. "Most of the time" we observe
behavior but do not know "the" function that describes it. We instead collect data about the behavior and make approximations
based off of this data. We demonstrate this in an example.

Example 5.5.8: Approximating distance traveled

One of the authors drove his daughter home from school while she recorded their speed every 30 seconds. The data is given in
Table 5.5.5. Approximate the distance they traveled.
Solution
Recall that by integrating a speed function we get distance traveled. We have information about v(t) ; we will use Simpson's
b
Rule to approximate ∫ v(t) dt.
a

The most difficult aspect of this problem is converting the given data into the form we need it to be in. The speed is measured
in miles per hour, whereas the time is measured in 30 second increments.
Table 5.5.5 : Speed data collected at 30 second intervals for Example 5.5.8 .
0 0

1 25

2 22

3 19

4 39

5.5.10 https://math.libretexts.org/@go/page/4182
5 0

6 43

7 59

8 54

9 51

10 43

11 35

12 40

13 43

14 30

15 0

16 0

17 28

18 40

19 42

20 40

21 39

22 40

23 23

24 0

We need to compute dx = (b − a)/n . Clearly, n = 24 . What are a and b ? Since we start at time t = 0 , we have that a =0 .
The final recorded time came after 24 periods of 30 seconds, which is 12 minutes or 1/5 of an hour. Thus we have
$$\ dx = \frac{b-a}{n} = \frac{1/5-0}{24} = \frac1{120}; \quad \frac{\ dx}{3} = \frac{1}{360}.\]
Thus the distance traveled is approximately:
0.2
1
∫ v(t) dt ≈ [f (x1 ) + 4f (x2 ) + 2f (x3 ) + ⋯ + 4f (xn ) + f (xn+1 )]
0
360

1
= [0 + 4 ⋅ 25 + 2 ⋅ 22 + ⋯ + 2 ⋅ 40 + 4 ⋅ 23 + 0]
360

≈ 6.2167 miles.

We approximate the author drove 6.2 miles. (Because we are sure the reader wants to know, the author's odometer recorded the
distance as about 6.05 miles.)

We started this chapter learning about antiderivatives and indefinite integrals. We then seemed to change focus by looking at areas
between the graph of a function and the x-axis. We defined these areas as the definite integral of the function, using a notation very
similar to the notation of the indefinite integral. The Fundamental Theorem of Calculus tied these two seemingly separate concepts
together: we can find areas under a curve, i.e., we can evaluate a definite integral, using antiderivatives.
We ended the chapter by noting that antiderivatives are sometimes more than difficult to find: they are impossible. Therefore we
developed numerical techniques that gave us good approximations of definite integrals.
We used the definite integral to compute areas, and also to compute displacements and distances traveled. There is far more we can
do than that. In Chapter 7 we'll see more applications of the definite integral. Before that, in Chapter 6 we'll learn advanced
techniques of integration, analogous to learning rules like the Product, Quotient and Chain Rules of differentiation.

5.5.11 https://math.libretexts.org/@go/page/4182
Contributors and Attributions
Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/
Integrated by Justin Marshall.

This page titled 5.5: Numerical Integration is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory
Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

5.5.12 https://math.libretexts.org/@go/page/4182
5.E: Applications of Integration (Exercises)
5.1: Antiderivatives and Indefinite Integration
Terms and Concepts
1. Define the term “antiderivative” in your own words
2. Is it more accurate to refer to “the” antiderivative of f (x) or “an” antiderivative of f (x)?
3. Use your own words to define the indefinite integral of f (x).
4. Fill in the blanks: “Inverse operations do the ____ things in the _____ order.”
5. What is an “initial value problem”?
6. The derivative of a position function is a _____ function.
7. The antiderivative of an acceleration function is a ______ function.

Problems
In Exercises 8-26, evaluate the given indefinite integral.
8. ∫ 3x 3
dx

9. ∫ x 8
dx

10. ∫ (10x 2
− 2) dx

11. ∫ dt

12. ∫ 1 ds
13. ∫ 3t
1
2
dt

14. ∫ 1
2
dt
t

15. ∫ √x
1
dx

16. ∫ sec 2
θ dθ

17. ∫ sin θ dθ
18. ∫ (sec x tan x + csc x cot x) dx
19. ∫ 5e θ

20. ∫ 3 t
dt

21. ∫ 5

2
dt

22. ∫ (2t + 3) 2
dt

23. ∫ (t 2
+ 3)(t
3
− 2t) dt

24. ∫ x 2
x
3
dx

25. ∫ e π
dx

26. ∫ a dx
27. This problem investigates why Theorem 35 states that ∫ dx = ln |x| + C .
1

(a) What is the domain of y = ln x ?


(b) Find (ln x). d

dx

(c) What is the domain of y = ln(−x) ?


(d) Find ((ln(−x)). d

dx

(e) You should find that 1/x has two types of antiderivatives, depending on whether x > 0 or x <0 . In one expression, give a
formula for ∫ dx that takes these different domains into account, and explain your answer.
1

5.E.1 https://math.libretexts.org/@go/page/9974
In Exercises 28-38, find f (x) described by the given initial value problem.
28. f ′
(x) = sin x and f (0) = 2

29. f ′
(x) = 5 e
x
and f (0) = 10

30. f ′
(x) = 4 x
3
− 3x
2
and f (−1) = 9

31. f ′
(x) = sec
2
x and f (π/4) = 5

32. f ′
(x) = 7
x
and f (2) = 1

33. f ′′ ′
(x) = 5 and f (0) = 7, f (0) = 3

34. f ′′ ′
(x) = 7x and f (1) = −1, f (1) = 10

35. f ′′
(x) = 5 e
x ′
and f (0) = 3, f (0) = 5

36. f ′′
(θ) = sin θ and f (π) = 2, f (π) = 4

37. f ′′
(x) = 24 x
2
+2
x ′
− cos x and f (0) = 5, f (0) = 0

38. f ′′ ′
(x) = 0 and f (1) = 3, f (1) = 1

Review
39. Use information gained from the first and second derivative to sketch f (x) = x
1

e +1
.
40. Given y = x 2
e
x
cos x , find dy .

5.2: The Definite Integral


Terms and Concepts
1. What is "total signed area"?
2. What is "displacement"?
3
3. What is ∫ 3
sin x dx

1 2
4. Give a single definite integral that has the same value as ∫ 0
(2x + 3) dx + ∫
1
(2x + 3) dx .

Problems
In Exercises 5-9, a graph of a function f (x) is given. Using the geometry of the graph, evaluate the definite integrals.
5.

1
(a) ∫0
(−2x + 4) dx
2
(b) ∫ 0
(−2x + 4) dx
3
(c) ∫0
(−2x + 4) dx
3
(d) ∫ 1
(−2x + 4) dx
4
(e) ∫2
(−2x + 4) dx
1
(f) ∫0
(−6x + 12) dx

5.E.2 https://math.libretexts.org/@go/page/9974
6.

2
(a) ∫ 0
f (x) dx
3
(b) ∫ 0
f (x) dx
5
(c) ∫ 0
f (x) dx
5
(d) ∫ 2
f (x) dx
3
(e) ∫ 5
f (x) dx
3
(f) ∫0
f (x) dx

7.

2
(a) ∫ 0
f (x) dx
4
(b) ∫ 2
f (x) dx
4
(c) ∫ 2
2f (x) dx
1
(d) ∫ 0
4x dx
3
(e) ∫ 2
(2x − 4) dx
3
(f) ∫2
(4x − 8) dx

8.

1
(a) ∫ 0
(x − 1) dx
2
(b) ∫ 0
(x − 1) dx
3
(c) ∫ 0
(x − 1) dx
3
(d) ∫ 2
(x − 1) dx
4
(e) ∫ 1
(x − 1) dx
4
(f) ∫1
((x − 1) + 1) dx

5.E.3 https://math.libretexts.org/@go/page/9974
9.

2
(a) ∫ 0
f (x) dx
4
(b) ∫ 2
f (x) dx
4
(c) ∫ 0
f (x) dx
4
(d) ∫ 0
5f (x) dx

In Exercises 10-13, 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.
10.

1
(a) ∫ 0
f (x) dx
2
(b) ∫ 0
f (x) dx
3
(c) ∫ 0
f (x) dx
2
(d) ∫ 1
−3f (x) dx

11.

2
(a) ∫ 0
f (x) dx
4
(b) ∫ 2
f (x) dx
4
(c) ∫ 0
f (x) dx
1
(d) ∫ 0
f (x) dx

12.

−1
(a) ∫ −2
f (x) dx

5.E.4 https://math.libretexts.org/@go/page/9974
2
(b) ∫ 1
f (x) dx
1
(c) ∫ −1
f (x) dx
1
(d) ∫ 0
f (x) dx

13.

2
(a) ∫ 0
5x
2
dx
2
(b) ∫ 0
(x
2
+ 1) dx
3
(c) ∫ 1
(x − 1 )
2
dx
4
(d) ∫ 2
((x − 2) + 5) dx

In Exercises 14-15, a graph of the velocity function of an object moving in a straight line is given. Answer the questions
based on that graph.
14.

(a) What is the object's maximum velocity?


(b) What is the object's maximum displacement?
(c) What is the object's total displacement on [0,3]?
15.

(a) What is the object's maximum velocity?


(b) What is the object's maximum displacement?
(c) What is the object's total displacement on [0,5]?
16. 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?
(b) What is the object's maximum displacement?

5.E.5 https://math.libretexts.org/@go/page/9974
(c) When does the maximum displacement occur?
(d) When will the object reach a height of 0? (Hint: find when the displacement is -48ft.)
17. An object is thrown straight up with a velocity, in ft/s, given by v(t) = −32t + 96 , where t is seconds, from a height of 64
feet.
(a) What is the object's initial velocity?
(b) What is the object's displacement 0?
(c) How long does it take for the object to return to its initial height?
(d) When will the object reach a height of 210ft?
In Exercises 18-21, let
2

0
f (x) dx = 5 ,
3

0
f (x) dx = 7 ,
2

0
g(x) dx = −3 , and
3

2
g(x) dx = 5 .
Use these values to evaluate the given definite integrals.
2
18. ∫ 0
(f (x) + g(x)) dx

3
19. ∫ 0
(f (x) − g(x)) dx

3
20. ∫ 2
(3f (x) + 2g(x)) dx

21. Find values for a and b such that


3
∫ (af (x) + bg(x)) dx = 0
0

In Exercises 22-25, let


3

0
s(t) dt = 10 ,
5

3
s(t) dt = 8 ,
5

3
r(t) dt = −1 , and
5

0
r(t) dt = 11 .
Use these values to evaluate the given definite integrals.
3
22. ∫ 0
(s(t) + r(t)) dt

0
23. ∫ 5
(s(t) − r(t)) dt

3
24. ∫ 3
(πs(t) − 7r(t)) dt

25. Find values for a and b such that


5
∫ (ar(t) + bs(t)) dt = 0
0

Review
In Exercises 26-29, evaluate the given indefinite integral.
26. ∫ (x 3
− 2x
2
+ 7x − 9) dx

27. ∫ (sin x − cos x + sec 2


x) dx

28. ∫ (√t + 3 1
2
t
t
+ 2 ) dt

29. ∫ ( 1

x
− csc x cot x) dx

5.3: Riemann Sums


Terms and Concepts
1. A fundamental calculus technique is to use ________ to refine approximations to get an exact answer.
14
2. What is the upper bound in the summation ∑ i=7
?
(48i − 201)

5.E.6 https://math.libretexts.org/@go/page/9974
3. This section approximates definite integrals using what geometric shape?
4. T/F: A sum using the Right Hand Rule is an example of a Riemann Sum.

Problems
In Exercises 5-11, write out each term of the summation and compute the sum.
5. ∑ 4

i=2
i
2

6. ∑ 3

i=−1
(4i − 2)

7. ∑ 2

i=−2
sin(πi/2)

8. ∑ 5

i=1
1

6
9. ∑ i=1
(−1 ) i
i

10. ∑ 4

i=1
(
1

i

i+1
1
)

11. ∑ 5

i=0
(−1 ) cos(πi)
i

In Exercises 12-15, write each sum in summation notation.


12. 3 + 6 + 9 + 12 + 15
13. −1 + 0 + 3 + 8 + 15 + 24 + 35 + 48 + 63
14. 1

2
+
2

3
+
3

4
+
4

15. 1 − e + e 2
−e
3
+e
4

In Exercises 16-22, evaluate the summation using Theorem 37.


16. ∑ 2

i=1
5i

1
17. ∑ i=1
0(3 i
2
− 2i)

1
18. ∑ i=1
5(2 i
3
− 10)

19. ∑ 1

i=1
0(−4 i
3
+ 10 i
2
− 7i + 11)

20. ∑ 1

i=1
0(i
3
− 3t
2
+ 2i + 7)

21. 1 + 2 + 3+. . . +99 + 100


22. 1 + 4 + 9+. . . +361 + 400
Theorem 37 states

n

i=1
ai = ∑ +i = 1
k
ai + ∑
n

i=k+1
ai , so

n

i=k+1
ai = ∑
n

i=1
ai − ∑
k

i=1
ai .
Use this fact, along with other parts of Theorem 37, to evaluate the summations given in Exercises 23-26.
2
23. ∑ i=11
0i

24. ∑ 2

i=16
5i
3

25. ∑ 1

i=7
24

26. ∑ 1

i=5
04 i
3

In Exercises 27-32, a definite integral


b
∫ f (x) dx is given.
a

(a) Graph f (x) on [a,b].


(b) Add to the sketch rectangles using the provided rule.
b
(c) Approximate ∫ f (x) dx by summing the areas of the rectangles.
a

5.E.7 https://math.libretexts.org/@go/page/9974
3
27. ∫ −3
x
2
dx , with 6 rectangles using the Left Hand Rule.
2
28. ∫ 0
(5 − x ) dx
2
, with 4 rectangles using the Midpoint Rule.
π
29. ∫ 0
sin x dx , with 6 rectangles using the Right Hand Rule.
3
30. ∫ 0
2
x
dx , with 5 rectangles using the Left Hand Rule.
2
31. ∫ 1
ln x dx , with 3 rectangles using the Midpoint Rule.
9
32. ∫ 1
1

x
dx , with 4 rectangles using the Right Hand Rule.
In Exercises 33-38, a definite integral
b
∫ f (x) dx is given. As demonstrated in Example 123 and 124, do the following.
a
b
(a) Find a formula to approximate ∫ f (x) dx using n subintervals and the provided rule.
a

(b) Evaluate the formula using n = 10, 100 and 1000.


b
(c) Find the limit of the formula, as n → ∞ , to find the exact value of ∫ a
.
f (x) dx

1
33. ∫ 0
x
3
dx , using the Right Hand Rule.
1
34. ∫ −1
3x
2
dx , using the Left Hand Rule.
3
35. ∫ −1
(3x − 1) dx , using the Midpoint Rule.
4
36. ∫ 1
(2 x
2
− 3) dx , using the Left Hand Rule.
10
37. ∫ −10
(5 − x) dx , using the Right Hand Rule.
1
38. ∫ 0
(x
3
− x ) dx
2
, using the Right Hand Rule.

Review
In Exercises 39-44, find an antiderivative of the given function.
39. f (x) = 5 sec 2
x

40. f (x) = 7

41. g(t) = 4t 5
− 5t
3
+8

42. g(t) = 5 ⋅ 8 t

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


44. f (x) = 1

√x

5.4: The Fundamental Theorem of Calculus


Terms and Concepts
1. How are the definite and indefinite integrals related?
2. What constants of integration is most commonly used when evaluating definite integrals?
x
3. T/F: If f is a continuous function, then F (x) = ∫ a
f (t) dt is also a continuous function.
4. The definite integral can be used to find "the area under a curve." Give two other uses for definite integrals.

Problems
In Exercises 5-28, evaluate the definite integral.
3
5. ∫ 1
(3 x
2
− 2x + 1) dx

4
6. ∫ 0
(x − 1 )
2
dx

1
7. ∫ −1
(x
3
− x ) dx
5

5.E.8 https://math.libretexts.org/@go/page/9974
π
8. ∫
π/2
cos x dx

π/4
9. ∫
0
sec
2
x dx

e
10. ∫ 1
1

x
dx

1
11. ∫ −1
5
x
dx

−1
12. ∫ −2
(4 − 2 x ) dx
3

π
13. ∫ 0
(2 cos x − 2 sin x) dx

3
14. ∫ 1
e
x
dx

4
15. ∫ 0
√t dt

25
16. ∫ 9
1

√t
dt

3 −
8
17. ∫ 1
√x dx

2
18. ∫ 1
1

x
dx

2
19. ∫ 1
1

x
2
dx

2
20. ∫ 1
1

x
3
dx

1
21. ∫ 0
x dx

1
22. ∫ 0
x
2
dx

1
23. ∫ 0
x
3
dx

1
24. ∫ 0
x
100
dx

4
25. ∫ −4
dx

−5
26. ∫ −10
3 dx

2
27. ∫ −2
0 dx

π/3
28. ∫ π/6
csc x cot x dx

29. Explain why:


1
(a) ∫ x dx = 0 , when n is a positive, odd integer, and
−1
n

1 1
(b) ∫ −1
x
n
dx = 2 ∫
0
n
x dx when n is a positive, even integer.
In Exercises 30-33, find a value c guaranteed by the Mean Value Theorem.
2
30. ∫ 0
x
2
dx

2
31. ∫ −2
x
2
dx

1
32. ∫ 0
e
x
dx

1 −
33 ∫ 0
6 √x dx

In Exercises 34-39, find the average value of the function on the given interval.
34. f (x) = sin x on [0, π/2]
35. y = sin x on [0, π]
36. y = x on [0, 4]
37. y = x 2
on [0, 4]

38. y = x 3
on [0, 4]

5.E.9 https://math.libretexts.org/@go/page/9974
39. g(t) = 1/t on [1, e]
In Exercises 40-44, a velocity function of an object moving along a straight line is given. Find the displacement of the object
over the given time interval.
40. v(t) = −32t + 20 ft/s on [0,5].
41. v(t) = −32t + 200 ft/s on [0,10].
42. v(t) = 2 mph on [-1,1].
t

43. v(t) = cos t ft/s on [0, 3π/2].


44. v(t) = √t ft/s on [0,16].
4

In Exercises 45-48, 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.
45. a(t) = −32 ft/s on [0,2].
46. a(t) = 10 ft/s on [0,5].
47. a(t) = t ft/s on [0,2]. 2

48. a(t) = cos t ft/s on [0, π]. 2

In Exercises 49-52, sketch the given functions and fine the area of the enclosed region.
49. y = 2x, y = 5x, and x = 3 .
50. y = −x + 1, y = 3x + 6, x = 2 and x = −1 .
51. y = x 2
− 2x + 5, y = 5x − 5 .
52. y = 2x 2
+ 2x − 5, y = x
2
+ 3x + 7 .
In Exercises 53-56, find F ′
(x) .
3
x +x
53. F (x) = ∫ 2
1

t
dt

0
54. F (x) = ∫ x3
t
3
dt

55. F (x) = x2
x
(t + 2) dt

x
e
56. F (x) = ∫ ln x
sin t dt

5.5: Numerical Integration


Terms and Concepts
1. T/F: Simpson's Rule is a method of approximating antiderivatives.
2. What are the two basic situations where approximating the value of a definite integral is necessary?
3. Why are the Left and Right Hand Rules rarely used?

Problems
In Exercises 4-11, a definite integral is given.
(a) Approximating the definite integral with the Trapezoidal Rule and n = 4 .
(b) Approximate the definite integral with Simpson's Rule and n = 4 .
(c) Find the exact value of the integral.
1
4. ∫
−1
x
2
dx

10
5. ∫
0
5x dx

π
6. ∫
0
sin x dx

4 −
7. ∫
0
√x dx

5.E.10 https://math.libretexts.org/@go/page/9974
3
8. ∫ 0
(x
3
+ 2x
2
− 5x + 7) dx

1
9. ∫ 0
x
4
dx

2x
10. ∫ 0
cos x dx

3 −−−− −
11. ∫ −3
√9 − x2 dx

In Exercises 12-19, approximate the definite integral with the Trapezoidal Rule and Simpson's Rule, with n = 6 .
1
12. ∫ 0
cos(x ) dx
2

2
1
13. ∫ −1
e
x
dx

5 −−−−−
14. ∫ 0
√x2 + 1 dx

π
15. ∫ 0
x sin x dx

π/2 −−−−
16. ∫ 0
√cos x dx

4
17. ∫ 1
ln x dx

1
18. ∫ −1 sin x+2
1
dx

6
19. ∫ 0 sin x+2
1
dx

In Exercises 20-23, find n such that the error in approximating the given definite integral is less than 0.0001 when using:
(a) the Trapezoidal Rule
(b) Simpson's Rule
π
20. ∫ 0
sin x dx

4
21. ∫ 1 √x
1
dx

π
22. ∫ 0
cos(x ) dx
2

5
23. ∫ 0
x
4
dx

In Exercises 24-25, a region is given. Find the area of the region using Simpson's Rule:
(a) where the measurements are in centimeters, taken in 1cm increments, and
(b) where the measurements are in hundreds of yards, taken in 100 yd increments.
24.

25.

5.E: Applications of Integration (Exercises) is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by LibreTexts.

5.E.11 https://math.libretexts.org/@go/page/9974
CHAPTER OVERVIEW
6: Techniques of Integration
6.1: Substitution
6.2: Integration by Parts
6.3: Trigonometric Integrals
6.4: Trigonometric Substitution
6.5: Partial Fraction Decomposition
6.6: Hyperbolic Functions
6.7: L'Hopital's Rule
6.8: Improper Integration
6.E: Applications of Antidifferentiation (Exercises)
Index

Thumbnail: Graph showing a few iterations of Newton's method on the graph y =x


2
with initial guess of x0 = 4 . (Public
Domain; Paul Breen).

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 6: Techniques of Integration is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory
Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

1
CHAPTER OVERVIEW
Front Matter
TitlePage
InfoPage

1
Virginia Military Institute
6: Techniques of Integration

Gregory Hartman et al.


This text is disseminated via the Open Education Resource (OER) LibreTexts Project (https://LibreTexts.org) and like the hundreds
of other texts available within this powerful platform, it is freely available for reading, printing and "consuming." Most, but not all,
pages in the library have licenses that may allow individuals to make changes, save, and print this book. Carefully
consult the applicable license(s) before pursuing such effects.
Instructors can adopt existing LibreTexts texts or Remix them to quickly build course-specific resources to meet the needs of their
students. Unlike traditional textbooks, LibreTexts’ web based origins allow powerful integration of advanced features and new
technologies to support learning.

The LibreTexts mission is to unite students, faculty and scholars in a cooperative effort to develop an easy-to-use online platform
for the construction, customization, and dissemination of OER content to reduce the burdens of unreasonable textbook costs to our
students and society. The LibreTexts project is a multi-institutional collaborative venture to develop the next generation of open-
access texts to improve postsecondary education at all levels of higher learning by developing an Open Access Resource
environment. The project currently consists of 14 independently operating and interconnected libraries that are constantly being
optimized by students, faculty, and outside experts to supplant conventional paper-based books. These free textbook alternatives are
organized within a central environment that is both vertically (from advance to basic level) and horizontally (across different fields)
integrated.
The LibreTexts libraries are Powered by MindTouch® and are supported by the Department of Education Open Textbook Pilot
Project, the UC Davis Office of the Provost, the UC Davis Library, the California State University Affordable Learning Solutions
Program, and Merlot. This material is based upon work supported by the National Science Foundation under Grant No. 1246120,
1525057, and 1413739. Unless otherwise noted, LibreTexts content is licensed by CC BY-NC-SA 3.0.
Any opinions, findings, and conclusions or recommendations expressed in this material are those of the author(s) and do not
necessarily reflect the views of the National Science Foundation nor the US Department of Education.
Have questions or comments? For information about adoptions or adaptions contact info@LibreTexts.org. More information on our
activities can be found via Facebook (https://facebook.com/Libretexts), Twitter (https://twitter.com/libretexts), or our blog
(http://Blog.Libretexts.org).

This text was compiled on 04/01/2023


6.1: Substitution
The previous chapter introduced the antiderivative and connected it to signed areas under a curve through the Fundamental
Theorem of Calculus. The next chapter explores more applications of definite integrals than just area. As evaluating definite
integrals will become important, we will want to find antiderivatives of a variety of functions.
This chapter is devoted to exploring techniques of antidifferentiation. While not every function has an antiderivative in terms of
elementary functions (a concept introduced in the section on Numerical Integration), we can still find antiderivatives of a wide
variety of functions.

Substitution
We motivate this section with an example. Let f (x) = (x 2
+ 3x − 5 )
10
. We can compute f ′
(x) using the Chain Rule. It is:
′ 2 9 2 9
f (x) = 10(x + 3x − 5 ) ⋅ (2x + 3) = (20x + 30)(x + 3x − 5 ) . (6.1.1)

Now consider this: What is ∫ (20x + 30)(x 2


+ 3x − 5 )
9
dx ? We have the answer in front of us;

2 9 2 10
∫ (20x + 30)(x + 3x − 5 ) dx = (x + 3x − 5 ) + C. (6.1.2)

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 knowing the original function first.
The underlying principle is to rewrite a "complicated" integral of the form ∫ f (x) dx as a not--so--complicated integral ∫ h(u) du .
We'll formally establish later how this is done. First, consider again our introductory indefinite integral,
∫ (20x + 30)(x + 3x − 5 ) dx . Arguably the most "complicated" part of the integrand is (x + 3x − 5 ) . We wish to make this
2 9 2 9

simpler; we do so through a substitution. Let u = x + 3x − 5 . Thus


2

2 9 9
(x + 3x − 5 ) =u . (6.1.3)

We have established u as a function of x, so now consider the differential of u:


du = (2x + 3)dx. (6.1.4)

Keep in mind that (2x + 3) and dx are multiplied; the dx is not "just sitting there."
Return to the original integral and do some substitutions through algebra:

2 9 2 9
∫ (20x + 30)(x + 3x − 5 ) dx =∫ 10(2x + 3)(x + 3x − 5 ) dx (6.1.5)

2 9
=∫ 10( x + 3x − 5 ) (2x + 3) dx (6.1.6)
 
u
du

9
=∫ 10 u du (6.1.7)

10 2
=u +C (replace u with x + 3x − 5) (6.1.8)

2 10
= (x + 3x − 5 ) +C (6.1.9)

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 derivative of their composition:
d ′ ′
(F (g(x))) = F (g(x))g (x). (6.1.10)
dx

Thus

6.1.1 https://math.libretexts.org/@go/page/4185
′ ′
∫ F (g(x))g (x) dx = F (g(x)) + C . (6.1.11)

Integration by substitution works by recognizing the "inside" function g(x) and replacing it with a variable. By setting u = g(x) ,
we can rewrite the derivative as
d
′ ′
(F (u)) = F (u)u . (6.1.12)
dx

Since du = g ′
(x)dx , we can rewrite the above integral as

′ ′ ′
∫ F (g(x))g (x) dx = ∫ F (u)du = F (u) + C = F (g(x)) + C . (6.1.13)

This concept is important so we restate it in the context of a theorem.

Theorem 6.1.1: Integration by Substitution

Let F and g be differentiable functions, where the range of g is an interval I contained in the domain of F . Then

′ ′
∫ F (g(x))g (x) dx = F (g(x)) + C . (6.1.14)

If u = g(x) , then du = g ′
(x)dx and

′ ′ ′
∫ F (g(x))g (x) dx = ∫ F (u) du = F (u) + C = F (g(x)) + C . (6.1.15)

The 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 derivative 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.

Example 6.1.1: Integrating by substitution

Evaluate ∫ x sin(x 2
+ 5) dx .
Solution
Knowing that substitution is related to the Chain Rule, we choose to let u be the "inside" function of sin(x 2
+ 5) . (This is not
always a good choice, but it is often the best place to start.)
Let u = x + 5 , hence du = 2x dx. The integrand has an x dx term, but not a 2x dx term. (Recall that multiplication is
2

commutative, so the x does not physically have to be next to dx for there to be an x dx term.) We can divide both sides of the
du expression by 2:

1
du = 2x dx ⇒ du = x dx. (6.1.16)
2

We can now substitute.

2 2
∫ x sin(x + 5) dx = ∫ sin( x + 5 )x dx (6.1.17)
 
1
u du
2

1
=∫ sin u du (6.1.18)
2

1
2
=− cos u + C (now replace u with x + 5) (6.1.19)
2

1 2
=− cos(x + 5) + C . (6.1.20)
2

Thus ∫ x sin(x 2
+ 5) dx = −
1

2
cos(x
2
+ 5) + C . We can check our work by evaluating the derivative of the right hand side.

6.1.2 https://math.libretexts.org/@go/page/4185
Example 6.1.2: Integrating by substitution

Evaluate ∫ cos(5x) dx.


Solution
Again let u replace the "inside" function. Letting u = 5x, we have du = 5dx . Since our integrand does not have a 5dx term,
we can divide the previous equation by 5 to obtain du = dx . We can now substitute.
1

∫ cos(5x) dx = ∫ cos( 5x ) dx (6.1.21)


 
u 1
du
5

1
=∫ cos u du (6.1.22)
5

1
= sin u + C (6.1.23)
5
1
= sin(5x) + C . (6.1.24)
5

We can again check our work through differentiation.

The previous example exhibited a common, and simple, type of substitution. The "inside" function was a linear function (in this
case, y = 5x). When the inside function is linear, the resulting integration is very predictable, outlined here.

Key Idea 10: Substitution With A Linear Function

Consider ∫ F ′
(ax + b) dx , where a ≠ 0 and b are constants. Letting u = ax + b gives du = a ⋅ dx , leading to the result


1
∫ F (ax + b) dx = F (ax + b) + C . (6.1.25)
a

Thus ∫ sin(7x − 4) dx = − 1

7
cos(7x − 4) + C . Our next example can use Key Idea 10, but we will only employ it after going
through all of the steps.

Example 6.1.3: Integrating by substituting a linear function

Evaluate ∫ 7

−3x+1
dx .
Solution
View this a composition of functions f (g(x)), where f (x) = 7/x and g(x) = −3x + 1 . Employing our understanding of
substitution, we let u = −3x + 1 , the inside function. Thus du = −3dx . The integrand lacks a −3; hence divide the previous
equation by −3 to obtain −du/3 = dx. We can now evaluate the integral through substitution.
7 7 du
∫ dx = ∫ (6.1.26)
−3x + 1 u −3

−7 du
= ∫ (6.1.27)
3 u

−7
= ln |u| + C (6.1.28)
3
7
=− ln | − 3x + 1| + C . (6.1.29)
3

Using Key Idea 10 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.

Not all integrals that benefit from substitution have a clear "inside" function. Several of the following examples will demonstrate
ways in which this occurs.

6.1.3 https://math.libretexts.org/@go/page/4185
Example 6.1.4: Integrating by substitution

Evaluate ∫ sin x cos x dx .


Solution
There is not a composition of function here to exploit; rather, just a product of functions. Do not be afraid to experiment; when
given an integral to evaluate, it is often beneficial to think "If I let u be this, then du must be that ..." and see if this helps
simplify the integral at all.
In this example, let's set u = sin x . Then du = cos x dx , which we have as part of the integrand! The substitution becomes
very straightforward:

∫ sin x cos x dx = ∫ u du (6.1.30)

1 2
= u +C (6.1.31)
2
1 2
= sin x + C. (6.1.32)
2

One would do well to ask "What would happen if we let u = cos x?" The result is just as easy to find, yet looks very different.
The challenge to the reader is to evaluate the integral letting u = cos x and discover why the answer is the same, yet looks
different.

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."

Example 6.1.5: Integrating by substitution


−−−−−
Evaluate ∫ x √x + 3 dx .
Solution
Recognizing the composition of functions, set u = x +3 . Then du = dx , giving what seems initially to be a simple
substitution. But at this stage, we have:

−− −−− −
∫ x √ x + 3 dx = ∫ x √u du. (6.1.33)

We cannot evaluate an integral that has both an x and an u in it. We need to convert the x to an expression involving just u.
Since we set u = x +3 , we can also state that u −3 = x . Thus we can replace x in the integrand with u −3 . It will also be
1

helpful to rewrite √u as u . 2

−− −−− 1

∫ x √ x + 3 dx = ∫ (u − 3)u 2 du (6.1.34)

3 1

=∫ (u 2
− 3u 2
) du (6.1.35)

2 5 3

= u 2
− 2u 2
+C (6.1.36)
5
2 5 3

= (x + 3 ) 2 − 2(x + 3 ) 2 + C. (6.1.37)
5

Checking your work is always a good idea. In this particular case, some algebra will be needed to make one's answer match the
integrand in the original problem.

Example 6.1.6: Integrating by substitution

Evaluate ∫ 1

x ln x
dx .
Solution

6.1.4 https://math.libretexts.org/@go/page/4185
This is another example where there does not seem to be an obvious composition of functions. The line of thinking used in
Example 6.1.5 is useful here: choose something for u and consider what this implies du must be. If u can be chosen such that
du also appears in the integrand, then we have chosen well.

Choosing u = 1/x makes du = −1/x 2


dx ; that does not seem helpful. However, setting u = ln x makes du = 1/x dx ,
which is part of the integrand. Thus:
1 1 1
∫ dx = ∫ dx (6.1.38)
x ln x ln x x
 
1/u du

1
=∫ du (6.1.39)
u

= ln |u| + C (6.1.40)

= ln | ln x| + C . (6.1.41)

The final answer is interesting; the natural log of the natural log. Take the derivative to confirm this answer is indeed correct.

Integrals Involving Trigonometric Functions


Section 6.3 delves deeper into integrals of a variety of trigonometric functions; here we use substitution to establish a foundation
that we will build upon.
The next three examples will help fill in some missing pieces of our antiderivative knowledge. We know the antiderivatives of the
sine and cosine functions; what about the other standard functions tangent, cotangent, secant and cosecant? We discover these next.

Example 6.1.7: Integration by substitution: antiderivatives of tan x

Evaluate ∫ tan x dx.


Solution
The previous paragraph established that we did not know the antiderivatives of tangent, hence we must assume that we have
learned something in this section that can help us evaluate this indefinite integral.
Rewrite tan x as sin x/ cos x. While the presence of a composition of functions may not be immediately obvious, recognize
that cos x is "inside" the 1/x function. Therefore, we see if setting u = cos x returns usable results. We have that
du = − sin x dx , hence −du = sin x dx . We can integrate:

sin x
∫ tan x dx = ∫ dx (6.1.42)
cos x

1
=∫ sin x dx (6.1.43)
cos x 
 −du
u

−1
=∫ du (6.1.44)
u

= − ln |u| + C (6.1.45)

= − ln | cos x| + C . (6.1.46)

Some texts prefer to bring the −1 inside the logarithm as a power of cos x, as in:
−1
− ln | cos x| + C = ln |(cos x ) | +C (6.1.47)

∣ 1 ∣
= ln∣ ∣+C (6.1.48)
∣ cos x ∣

= ln | sec x| + C . (6.1.49)

Thus the result they give is ∫ tan x dx = ln | sec x| + C . These two answers are equivalent.

6.1.5 https://math.libretexts.org/@go/page/4185
Example 6.1.8: Integrating by substitution: antiderivatives of sec x

Evaluate ∫ sec x dx .
Solution
This example employs a wonderful trick: multiply the integrand by "1" so that we see how to integrate more clearly. In this
case, we write "1" as
sec x + tan x
1 = . (6.1.50)
sec x + tan x

This may seem like it came out of left field, but it works beautifully. Consider:
sec x + tan x
∫ sec x dx = ∫ sec x ⋅ dx (6.1.51)
sec x + tan x
2
sec x + sec x tan x
=∫ dx. (6.1.52)
sec x + tan x

Now let u = sec x + tan x ; this means du = (sec x tan x + sec 2


x) dx , which is our numerator. Thus:
du
=∫ (6.1.53)
u

= ln |u| + C (6.1.54)

= ln | sec x + tan x| + C . (6.1.55)

We can use similar techniques to those used in Examples 6.1.6 and 6.1.7 to find antiderivatives of cot x and csc x (which the
reader can explore in the exercises.) We summarize our results here.

Theorem 6.1.1: Antiderivatives of Trigonometric Functions

1. ∫ sin x dx = − cos x + C
2. ∫ cos x dx = sin x + C
3. ∫ tan x dx = − ln | cos x| + C
4. ∫ csc x dx = − ln | csc x + cot x| + C
5. ∫ sec x dx = ln | sec x + tan x| + C
6. ∫ cot x dx = ln | sin x| + C

We explore one more common trigonometric integral.

Example 6.1.9: Integration by substitution: powers of cos x and sin x

Evaluate ∫ cos 2
x dx .
Solution
2
We have a composition of functions as cos 2
x = ( cos x ) .
However, setting u = cos x means du = − sin x dx , which we do not have in the integral. Another technique is needed.
The process we'll employ is to use a Power Reducing formula for cos 2
x (perhaps consult the back of this text for this formula),
which states
1 + cos(2x)
2
cos x = . (6.1.56)
2

The right hand side of this equation is not difficult to integrate. We have:
1 + cos(2x)
2
∫ cos x dx = ∫ dx (6.1.57)
2

1 1
=∫ ( + cos(2x)) dx. (6.1.58)
2 2

6.1.6 https://math.libretexts.org/@go/page/4185
Now use Key Idea 10:
1 1 sin(2x)
= x+ +C (6.1.59)
2 2 2

1 sin(2x)
= x+ + C. (6.1.60)
2 4

We'll make significant use of this power--reducing technique in future sections.

Simplifying the Integrand


It is common to be reluctant to manipulate the integrand of an integral; at first, our grasp of integration is tenuous and one may
think that working with the integrand will improperly change the results. Integration by substitution works using a different logic:
as long as equality is maintained, the integrand can be manipulated so that its form is easier to deal with. The next two examples
demonstrate common ways in which using algebra first makes the integration easier to perform.

Example 6.1.10: Integration by substitution: simplifying first


3 2
x + 4x + 8x + 5
Evaluate ∫ 2
dx .
x + 2x + 1

Solution
One may try to start by setting u equal to either the numerator or denominator; in each instance, the result is not workable.
When dealing with rational functions (i.e., quotients made up of polynomial functions), it is an almost universal rule that
everything works better when the degree of the numerator is less than the degree of the denominator. Hence we use polynomial
division.
We skip the specifics of the steps, but note that when x 2
+ 2x + 1 is divided into x 3
+ 4x
2
+ 8x + 5 , it goes in x +2 times
with a remainder of 3x + 3 . Thus
3 2
x + 4x + 8x + 5 3x + 3
= x +2 + . (6.1.61)
2 2
x + 2x + 1 x + 2x + 1

Integrating x + 2 is simple. The fraction can be integrated by setting u = x + 2x + 1 , giving du = (2x + 2) dx . This is very
2

similar to the numerator. Note that du/2 = (x + 1) dx and then consider the following:
3 2
x + 4x + 8x + 5 3x + 3
∫ dx = ∫ (x + 2 + ) dx (6.1.62)
2 2
x + 2x + 1 x + 2x + 1

3(x + 1)
= ∫ (x + 2) dx + ∫ dx (6.1.63)
2
x + 2x + 1

1 2
3 du
= x + 2x + C1 + ∫ (6.1.64)
2 u 2

1 3
2
= x + 2x + C1 + ln |u| + C2 (6.1.65)
2 2
1 3
2 2
= x + 2x + ln | x + 2x + 1| + C . (6.1.66)
2 2

In some ways, we "lucked out" in that after dividing, substitution was able to be done. In later sections we'll develop
techniques for handling rational functions where substitution is not directly feasible.

Example 6.1.11: Integration by alternate methods


2
x + 2x + 3
Evaluate ∫ − dx with, and without, substitution.
√x

Solution
1

We already know how to integrate this particular example. Rewrite √−


x as x and simplify the fraction: 2

6.1.7 https://math.libretexts.org/@go/page/4185
2
x + 2x + 3 3 1

1

=x 2
+ 2x 2
+ 3x 2
. (6.1.67)
1/2
x

We can now integrate using the Power Rule:


2
x + 2x + 3 3 1 1

∫ dx = ∫ (x 2 + 2x 2 + 3x 2 ) dx (6.1.68)
1/2
x

2 5 4 3 1

= x 2
+ x 2
+ 6x 2
+C (6.1.69)
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 ; therefore2

\[du = \frac12x^{-\frac12}dx = \frac{1}{2\sqrt{x}}\ dx \quad \Rightarrow \quad 2du = \frac{1}{\sqrt{x}}\ dx.$$

2
x + 2x + 3 1

This gives us ∫ − dx = ∫ (x
2
+ 2x + 3) ⋅ 2 du . What are we to do with the other x terms? Since u =x 2
,
√x

u
2
=x , etc. We can then replace x and x with appropriate powers of u. We thus have
2

2
x + 2x + 3
2
∫ − dx = ∫ (x + 2x + 3) ⋅ 2 du
√x

4 2
=∫ 2(u + 2u + 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.

Substitution and Inverse Trigonometric Functions


When studying derivatives of inverse functions, we learned that
d 1
−1
( tan x) = . (6.1.70)
2
dx 1 +x

Applying the Chain Rule to this is not difficult; for instance,


d 5
−1
( tan 5x) = . (6.1.71)
2
dx 1 + 25x

We now explore how Substitution can be used to "undo" certain derivatives that are the result of the Chain Rule applied to Inverse
Trigonometric functions. We begin with an example.

Example 6.1.12: Integrating by substitution: inverse trigonometric functions

Evaluate ∫ 1

25+x2
dx .

Solution
The integrand looks similar to the derivative of the arctangent function. Note:

6.1.8 https://math.libretexts.org/@go/page/4185
1 1
= (6.1.72)
x2
25 + x2 25(1 + )
25

1
= (6.1.73)
x 2
25(1 + ( ) )
5

1 1
= . (6.1.74)
2
25 x
1 +( )
5

Thus
1 1 1
∫ dx = ∫ dx. (6.1.75)
2 2
25 + x 25 x
1 +( )
5

This can be integrated using Substitution. Set u = x/5, hence du = dx/5 or dx = 5du . Thus
1 1 1
∫ dx = ∫ dx (6.1.76)
2 2
25 + x 25 x
1 +( )
5

1 1
= ∫ du (6.1.77)
2
5 1 +u

1 −1
= tan u +C (6.1.78)
5
1 −1
x
= tan ( )+C (6.1.79)
5 5

Example 6.1.12 demonstrates a general technique that can be applied to other integrands that result in inverse trigonometric
functions. The results are summarized here.

Theorem 6.1.2: Integrals Involving Inverse Trigonomentric Functions

Let a > 0 .
1 1 x
1. ∫ 2 2
dx =
−1
tan ( )+C
a +x a a
1 x
2. ∫ −−−− −−
dx = sin
−1
( )+C
√a2 − x2 a

1 1 |x|
3. ∫ −− −−−− dx = sec
−1
( ) +C
2
x √x − a
2 a a

Let's practice using Theorem 6.1.2.

Example 6.1.13: Integrating by substitution: inverse trigonometric functions

Evaluate the given indefinite integrals.


$$\displaystyle \int \frac{1}{9+x^2}\ dx,\quad \int \frac{1}{x\sqrt{x^2-\frac{1}{100}}}\ dx\quad \text{ and }\quad \int
\frac{1}{\sqrt{5-x^2}}\ dx.\]
Solution
Each can be answered using a straightforward application of Theorem 6.1.2.
1 1 x

2
dx = tan
−1
+C , as a = 3 .
9 +x 3 3

1

−−−−−−−
dx = 10 sec
−1
10x + C , as a = 1

10
.
2 1
x √x −
100

1 x –
∫ −−−− − = sin
−1

+C , as a = √5 .
√5 − x2 √5

6.1.9 https://math.libretexts.org/@go/page/4185
Most applications of Theorem 6.1.2 are not as straightforward. The next examples show some common integrals that can still be
approached with this theorem.

Example 6.1.14: Integrating by substitution: completing the square


1
Evaluate ∫ 2
dx .
x − 4x + 13

Solution
Initially, this integral seems to have nothing in common with the integrals in Theorem 6.1.2. As it lacks a square root, it almost
certainly is not related to arcsine or arcsecant. It is, however, related to the arctangent function.
We see this by completing the square in the denominator. We give a brief reminder of the process here.
Start with a quadratic with a leading coefficient of 1. It will have the form of x
2
+ bx + c . Take 1/2 of b , square it, and
add/subtract it back into the expression. I.e.,
2 2
b b
2 2
x + bx + c = x + bx + − +c (6.1.80)
4 4

2
(x+b/2)

2 2
b b
= (x + ) +c − (6.1.81)
2 4

In our example, we take half of −4 and square it, getting 4. We add/subtract it into the denominator as follows:
1 1
= (6.1.82)
2 2
x − 4x + 13 x − 4x + 4 − 4 + 13

2
(x−2)

1
= (6.1.83)
2
(x − 2 ) +9

We can now integrate this using the arctangent rule. Technically, we need to substitute first with u = x − 2 , but we can employ
Key Idea 10 instead. Thus we have
1 1 1 −1
x −2
∫ dx = ∫ dx = tan + C. (6.1.84)
2 2
x − 4x + 13 (x − 2 ) +9 3 3

Example 6.1.15: Integrals requiring multiple methods


4 −x
Evaluate ∫ −−−−− −
dx .
√16 − x2

Solution
This integral requires two different methods to evaluate it. We get to those methods by splitting up the integral:
4 −x 4 x
∫ − −−−− − dx = ∫ − −−−− − dx − ∫ − −−−− − dx. (6.1.85)
√ 16 − x2 √ 16 − x2 √ 16 − x2

The first integral is handled using a straightforward application of Theorem 6.1.2 ; the second integral is handled by
substitution, with u = 16 − x . We handle each separately.
2

4 x
−1
∫ dx = 4 sin + C.
−−−−− −
√16 − x2 4

x

−−−−− −
dx : Set u = 16 − x , so du = −2xdx and xdx = −du/2. We have
2

√16 − x2

6.1.10 https://math.libretexts.org/@go/page/4185
x −du/2
∫ dx = ∫ (6.1.86)
− −−−− − −
√ 16 − x2 √u

1 1
=− ∫ − du (6.1.87)
2 √u

= −√u + C (6.1.88)
− −−−− −
2
= −√ 16 − x + C . (6.1.89)

Combining these together, we have


4 −x x − −−−− −
−1 2
∫ − −−−− − dx = 4 sin + √ 16 − x + C . (6.1.90)
√ 16 − x2 4

Substitution and Definite Integration


This section has focused on evaluating indefinite integrals as we are learning a new technique for finding antiderivatives. However,
much of the time integration is used in the context of a definite integral. Definite integrals that require substitution can be
calculated using the following workflow:
b

1. Start with a definite integral ∫ f (x) dx that requires substitution.


a

2. Ignore the bounds; use substitution to evaluate ∫ f (x) dx and find an antiderivative F (x).
b

3. Evaluate F (x) at the bounds; that is, evaluate F (x)∣∣ = F (b) − F (a) .
a

This workflow works fine, but substitution offers an alternative that is powerful and amazing (and a little time saving).

At its heart, (using the notation of Theorem 6.1.1 substitution converts integrals of the form ∫ ′ ′
F (g(x))g (x) dx into an integral

of the form ∫ ′
F (u) du with the substitution of u = g(x) . The following theorem states how the bounds of a definite integral can
be changed as the substitution is performed.

Theorem 6.1.3: Substitution with Definite Integrals

Let F and g be differentiable functions, where the range of g is an interval I that is contained in the domain of F . Then
b g(b)
′ ′ ′
∫ F (g(x))g (x) dx = ∫ F (u) du. (6.1.91)
a g(a)

In effect, Theorem 6.1.3 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.

Example 6.1.16: Definite integrals and substitution: changing the bounds


2

Evaluate ∫ cos(3x − 1) dx using Theorem 6.1.3.


0

Solution
Observing the composition of functions, let u = 3x − 1 , hence du = 3dx. As\(3dx\) does not appear in the integrand, divide
the latter equation by 3 to get du/3 = dx .
By setting u = 3x − 1 , we are implicitly stating that g(x) = 3x − 1 . Theorem 6.1.3 states that the new lower bound is

g(0) = −1 ; the new upper bound is g(2) = 5 . We now evaluate the definite integral:

6.1.11 https://math.libretexts.org/@go/page/4185
2 5
du
∫ cos(3x − 1) dx = ∫ cos u (6.1.92)
1 −1 3

1 5

= sin u (6.1.93)

3 −1

1
= ( sin 5 − sin(−1)) ≈ −0.039. (6.1.94)
3

Notice how once we converted the integral to be in terms of u, we never went back to using x.

Figure 6.1.1 : Graphing the areas defined by the definite integrals of Example 6.1.16
The graphs in Figure 6.1.1 tell more of the story. In (a) the area defined by the original integrand is shaded, whereas in (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.

Example 6.1.17: Definite integrals and substitution: changing the bounds


π/2

Evaluate ∫ sin x cos x dx using Theorem 6.1.3.


0

Solution
We saw the corresponding indefinite integral in Example 6.1.4. In that example we set u = sin x but stated that we could have
let u = cos x. For variety, we do the latter here.
Let u = g(x) = cos x , giving du = − sin x dx and hence sin x dx = −du . The new upper bound is g(π/2) = 0 ; the new
lower bound is g(0) = 1 . Note how the lower bound is actually larger than the upper bound now. We have
π/2 0

∫ sin x cos x dx = ∫ −u du \scriptsize (switch bounds \& change sign) (6.1.95)


0 1

=∫ u du (6.1.96)
0

1 1
2∣
= u = 1/2. (6.1.97)
∣0
2

In Figure 6.1.2 we have again graphed the two regions defined by our definite integrals. Unlike the previous example, they
bear no resemblance to each other. However, Theorem 6.1.3 guarantees that they have the same area.

6.1.12 https://math.libretexts.org/@go/page/4185
Figure 6.1.2 : Graphing the areas defined by the definite integrals of Example 6.1.17.

Integration by substitution is a powerful and useful integration technique. The next section introduces another technique, called
Integration by Parts. As substitution "undoes" the Chain Rule, integration by parts "undoes" the Product Rule. Together, these two
techniques provide a strong foundation on which most other integration techniques are based.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/
Integrated by Justin Marshall.

This page titled 6.1: Substitution is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et al.
via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

6.1.13 https://math.libretexts.org/@go/page/4185
6.2: Integration by Parts
Here's a simple integral that we can't yet evaluate:
$$\int x\cos x \,dx.\]
It's a simple matter to take the derivative of the integrand using the Product Rule, but there is no Product Rule for integrals.
However, this section introduces Integration by Parts, a method of integration that is based on the Product Rule for derivatives. It
will enable us to evaluate this integral.
The Product Rule says that if u and v are functions of x, then (uv) = u v + u v . For simplicity, we've written
′ ′ ′
u for u(x) and v

for v(x). Suppose we integrate both sides with respect to x. This gives
$$\int (uv)'\,dx = \int (u'v+uv')\,dx.\]
By the Fundamental Theorem of Calculus, the left side integrates to uv. The right side can be broken up into two integrals, and we
have
$$uv = \int u'v\,dx + \int uv'\,dx.\]
Solving for the second integral we have
$$\int uv'\,dx = uv - \int u'v\,dx.\]
Using differential notation, we can write du = u (x)dx and dv = v (x)dx and the expression above can be written as follows:
′ ′

$$\int u\,dv = uv - \int v\,du.\]


This is the Integration by Parts formula. For reference purposes, we state this in a theorem.

Theorem 6.2.1: Integration by Parts

Let u and v be differentiable functions of x on an interval I containing a and b . Then

∫ u dv = uv − ∫ v du, (6.2.1)

and integration by parts


x=b x=b
b

∫ u dv = uv −∫ v du. (6.2.2)

a
x=a x=a

Let's try an example to understand our new technique.

Example 6.2.1: Integrating using Integration by Parts

Evaluate ∫ x cos x dx .

Solution
The key to Integration by Parts is to identify part of the integrand as "u" and part as "dv ." Regular practice will help one make
good identifications, and later we will introduce some principles that help. For now, let u = x and dv = cos x dx .
It is generally useful to make a small table of these values as done below. Right now we only know u and dv as shown on the
left of Figure 6.2.1; on the right we fill in the rest of what we need. If u = x , then du = dx . Since dv = cos x dx , v is an
antiderivative of cos x. We choose v = sin x .

Figure 6.2.1 : Setting up integration by parts.


Now substitute all of this into the Integration by Parts formula, giving
$$\int x\cos x\,dx = x\sin x - \int \sin x \,dx.\]

6.2.1 https://math.libretexts.org/@go/page/4186
We can then integrate sin x to get − cos x + C and overall our answer is
$$\int x\cos x\ dx = x\sin x + \cos x + C.\]
Note how the antiderivative contains a product, x sin x. This product is what makes Integration by Parts necessary.

The example above demonstrates how Integration by Parts works in general. We try to identify u and dv in the integral we are
given, and the key is that we usually want to choose u and dv so that du is simpler than u and v is hopefully not too much more
complicated than dv . This will mean that the integral on the right side of the Integration by Parts formula, ∫ v du will be simpler to
integrate than the original integral ∫ u dv .
In the example above, we chose u = x and dv = cos x dx . Then du = dx was simpler than u and v = sin x is no more
complicated than dv . Therefore, instead of integrating x cos x dx, we could integrate sin x dx, which we knew how to do.
A useful mnemonic for helping to determine u is "LIATE," where

L = Logarithmic, I = InverseT rig. , A = Algebraic(polynomials), (6.2.3)

T = Trigonometric, andE = Exponential. (6.2.4)

If the integrand contains both a logarithmic and an algebraic term, in general letting u be the logarithmic term works best, as
indicated by L coming before A in LIATE.
We now consider another example.

Example 6.2.2: Integrating using Integration by Parts

Evaluate ∫ xe
x
dx .

Solution
The integrand contains an Algebraic term (x) and an \textbf{E}xponential term (e ). Our mnemonic suggests letting u be the
x

algebraic term, so we choose u = x and dv = e dx . Then du = dx and v = e as indicated by the tables below.
x x

Figure 6.2.2 : Setting up Integration by Parts.


We see du is simpler than u, while there is no change in going from dv to v . This is good. The Integration by Parts formula
gives
$$\int x e^x\,dx = xe^x - \int e^x\,dx.\]
The integral on the right is simple; our final answer is
$$\int xe^x\ dx = xe^x - e^x + C.\]
Note again how the antiderivatives contain a product term.

Example 6.2.3: Integrating using Integration by Parts

Evaluate ∫ x
2
cos x dx .

Solution
The mnemonic suggests letting u =x
2
instead of the trigonometric function, hence dv = cos x dx . Then du = 2x dx and
v = sin x as shown below.

Figure 6.2.3 : Setting up Integration by Parts.

6.2.2 https://math.libretexts.org/@go/page/4186
The Integration by Parts formula gives
$$\int x^2\cos x\,dx = x^2\sin x - \int 2x\sin x\,dx.\]
At this point, the integral on the right is indeed simpler than the one we started with, but to evaluate it, we need to do
Integration by Parts again. Here we choose u = 2x and dv = sin x and fill in the rest below.

Figure 6.2.4 : Setting up Integration by Parts.


The integral all the way on the right is now something we can evaluate. It evaluates to −2 sin x . Then going through and
simplifying, being careful to keep all the signs straight, our answer is
$$\int x^2\cos x\ dx = x^2\sin x + 2x\cos x - 2\sin x + C.\]

Example 6.2.4: Integrating using Integration by Parts

Evaluate ∫ e
x
.
cos x dx

Solution
This is a classic problem. Our mnemonic suggests letting u be the trigonometric function instead of the exponential. In this
particular example, one can let u be either cos x or e ; to demonstrate that we do not have to follow LIATE, we choose u = e
x x

and hence dv = cos x dx . Then du = e dx and v = sin x as shown below.


x

Figure 6.2.5 : Setting up Integration by Parts.


Notice that du is no simpler than u, going against our general rule (but bear with us). The Integration by Parts formula yields
$$\int e^x\cos x\ dx = e^x\sin x - \int e^x\sin x\,dx.\]
The integral on the right is not much different than the one we started with, so it seems like we have gotten nowhere. Let's keep
working and apply Integration by Parts to the new integral, using u = e and dv = sin x dx . This leads us to the following:
x

Figure 6.2.6 : Setting up Integration by Parts.


The Integration by Parts formula then gives:

x x x x
∫ e cos x dx =e sin x − (−e cos x − ∫ −e cos x dx)

x x x
=e sin x + e cos x − ∫ e cos x dx.

It seems we are back right where we started, as the right hand side contains ∫ e x
cos x dx . But this is actually a good thing.
Add ∫ e x
cos x dx to both sides. This gives

x x x
2∫ e cos x dx = e sin x + e cos x

1
x x x
Now divide both sides by 2: ∫ e cos x dx = (e sin x + e cos x).
2

Simplifying a little and adding the constant of integration, our answer is thus
$$\int e^x\cos x\ dx = \frac12e^x\left(\sin x + \cos x\right)+C.\]

6.2.3 https://math.libretexts.org/@go/page/4186
Example 6.2.5: Integrating using Integration by Parts: antiderivative of ln x

Evaluate ∫ ln x dx .

Solution
One may have noticed that we have rules for integrating the familiar trigonometric functions and e , but we have not yet given
x

a rule for integrating ln x. That is because ln x can't easily be integrated with any of the rules we have learned up to this point.
But we can find its antiderivative by a clever application of Integration by Parts. Set u = ln x and dv = dx . This is a good,
sneaky trick to learn as it can help in other situations. This determines du = (1/x) dx and v = x as shown below.

Figure 6.2.7 : Setting up Integration by Parts.


Putting this all together in the Integration by Parts formula, things work out very nicely:
$$\int \ln x\,dx = x\ln x - \int x\,\frac1x\,dx.\]
The new integral simplifies to ∫ 1 dx , which is about as simple as things get. Its integral is x + C and our answer is
$$\int \ln x\ dx = x\ln{x} - x + C.\]

Example 6.2.6: Integrating using Int. by Parts: antiderivative of arctan x

Evaluate ∫ arctan x dx .

Solution
The same sneaky trick we used above works here. Let u = arctan x and dv = dx . Then du = 1/(1 + x 2
) dx and v = x . The
Integration by Parts formula gives
x
∫ arctan x dx = x arctan x − ∫ dx. (6.2.5)
2
1 +x

The integral on the right can be solved by substitution. Taking u = 1 + x , we get du = 2x dx. The integral then becomes
2

1 1
∫ arctan x dx = x arctan x − ∫ du. (6.2.6)
2 u

The integral on the right evaluates to ln |u| + C , which becomes ln(1 + x 2


)+C . Therefore, the answer is

2
∫ arctan x dx = x arctan x − ln(1 + x ) + C . (6.2.7)

Substitution Before Integration


When taking derivatives, it was common to employ multiple rules (such as using both the Quotient and the Chain Rules). It should
then come as no surprise that some integrals are best evaluated by combining integration techniques. In particular, here we illustrate
making an "unusual" substitution first before using Integration by Parts.

Example 6.2.7: Integration by Parts after substitution

Evaluate ∫ cos(ln x) dx .

Solution
The integrand contains a composition of functions, leading us to think Substitution would be beneficial. Letting u = ln x , we
have du = 1/x dx. This seems problematic, as we do not have a 1/x in the integrand. But consider:

6.2.4 https://math.libretexts.org/@go/page/4186
1
du = dx ⇒ x ⋅ du = dx. (6.2.8)
x

Since u = ln x , we can use inverse functions and conclude that x = e . Therefore we have that u

dx = x ⋅ du
u
=e du.

We can thus replace ln x with u and dx with e u


du . Thus we rewrite our integral as

u
∫ cos(ln x) dx = ∫ e cos u du. (6.2.9)

We evaluated this integral in Example 6.2.4. Using the result there, we have:

u
∫ cos(ln x) dx = ∫ e cos u du

1
u
= e ( sin u + cos u) + C
2
1
ln x
= e ( sin(ln x) + cos(ln x)) + C
2
1
= x( sin(ln x) + cos(ln x)) + C .
2

Definite Integrals and Integration By Parts


So far we have focused only on evaluating indefinite integrals. Of course, we can use Integration by Parts to evaluate definite
integrals as well, as Theorem 6.2.1 states. We do so in the next example.

Example 6.2.8: Definite integration using Integration by Parts


2

Evaluate ∫ 2
x .
ln x dx
1

Solution
Our mnemonic suggests letting u = ln x , hence dv = x 2
dx .
We then get du = (1/x) dx and v = x 3
/3 as shown below.

Figure 6.2.8 : Setting up Integration by Parts.


The Integration by Parts formula then gives
2 3 2 2 3
x ∣ x 1
2
∫ x ln x dx = ln x ∣ − ∫ dx
3 ∣ 3 x
1 1 1

3 2 2 2
x ∣ x
= ln x ∣ − ∫ dx
3 ∣1 3
1

3 2 3 2
x ∣ x ∣
= ln x ∣ − ∣
3 ∣1 9 ∣1
3 3 2
x x ∣
=( ln x − )∣
3 9 ∣1

8 8 1 1
=( ln 2 − ) −( ln 1 − )
3 9 3 9

8 7
= ln 2 −
3 9

≈ 1.07.

6.2.5 https://math.libretexts.org/@go/page/4186
In general, Integration by Parts is useful for integrating certain products of functions, like ∫ xe
x
dx or 3
∫ x sin x dx . It is also
useful for integrals involving logarithms and inverse trigonometric functions.
As stated before, integration is generally more difficult than derivation. We are developing tools for handling a large array of
integrals, and experience will tell us when one tool is preferable/necessary over another. For instance, consider the three similar--
looking integrals
$$\int xe^x\,dx, \qquad \int x e^{x^2}\,dx \qquad \text{and} \qquad \int xe^{x^3}\,dx.\]
While the first is calculated easily with Integration by Parts, the second is best approached with Substitution. Taking things one step
further, the third integral has no answer in terms of elementary functions, so none of the methods we learn in calculus will get us
the exact answer.
Integration by Parts is a very useful method, second only to substitution. In the following sections of this chapter, we continue to
learn other integration techniques. The next section focuses on handling integrals containing trigonometric functions.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/
Integrated by Justin Marshall.

This page titled 6.2: Integration by Parts is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman
et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

6.2.6 https://math.libretexts.org/@go/page/4186
6.3: Trigonometric Integrals
Functions involving trigonometric functions are useful as they are good at describing periodic behavior. This section describes
several techniques for finding antiderivatives of certain combinations of trigonometric functions.

Integrals of the form ∫ sin m


x cos
n
x dx

In learning the technique of Substitution, we saw the integral ∫ sin x cos x dx in Example 6.1.4. The integration was not difficult,
and one could easily evaluate the indefinite integral by letting u = sin x or by letting u = cos x. This integral is easy since the
power of both sine and cosine is 1.
We generalize this integral and consider integrals of the form ∫ sin x cos x dx , where m, n are nonnegative integers. Our m n

strategy for evaluating these integrals is to use the identity cos x + sin x = 1 to convert high powers of one trigonometric
2 2

function into the other, leaving a single sine or cosine term in the integrand. We summarize the general technique in the following
Key Idea.

Key Idea 11: Integrals Involving Powers of Sine and Cosine


Consider ∫ sin m
x cos
n
x dx , where m, n are nonnegative integers.
1. If m is odd, then m = 2k + 1 for some integer k . Rewrite
m 2k+1 2k 2 k 2 k
sin x = sin x = sin x sin x = (sin x) sin x = (1 − cos x) sin x. (6.3.1)

Then

m n 2 k n 2 k n
∫ sin x cos x dx = ∫ (1 − cos x) sin x cos x dx = − ∫ (1 − u ) u du, (6.3.2)

where u = cos x and du = − sin x dx .


2. If n is odd, then using substitutions similar to that outlined above we have

m n m 2 k
∫ sin x cos x dx = ∫ u (1 − u ) du, (6.3.3)

where u = sin x and du = cos x dx .


3. If both m and n are even, use the power--reducing identities
1 + cos(2x) 1 − cos(2x)
2 2
cos x = and sin x = (6.3.4)
2 2

to reduce the degree of the integrand. Expand the result and apply the principles of this Key Idea again.

We practice applying Key Idea 11 in the next examples.

Example 6.3.1: Integrating powers of sine and cosine

Evaluate ∫ sin 5
x cos
8
x dx .
Solution
The power of the sine term is odd, so we rewrite sin 5
x as
5 4 2 2 2 2
sin x = sin x sin x = (sin x) sin x = (1 − cos x) sin x.

Our integral is now ∫ (1 − cos 2


x)
2
cos
8
x sin x dx . Let u = cos x , hence du = − sin x dx . Making the substitution and
expanding the integrand gives

2 2 8 2 2 8 2 4 8 8 10 12
∫ (1 − cos ) cos x sin x dx = − ∫ (1 − u ) u du = − ∫ (1 − 2 u + u )u du = − ∫ (u − 2u +u ) du.

This final integral is not difficult to evaluate, giving

6.3.1 https://math.libretexts.org/@go/page/4187
8 10 12
1 9
2 11
1 13
−∫ (u − 2u +u ) du =− u + u − u +C
9 11 13

1 2 1
9 11 13
=− cos x+ cos x− cos x + C.
9 11 13

Example 6.3.2: Integrating powers of sine and cosine

Evaluate ∫ sin5
x cos
9
x dx .
Solution
The powers of both the sine and cosine terms are odd, therefore we can apply the techniques of Key Idea 11 to either power.
We choose to work with the power of the cosine term since the previous example used the sine term's power.
We rewrite cos 9
x as
9 8
cos x = cos x cos x

2 4
= (cos x) cos x

2 4
= (1 − sin x) cos x

2 4 6 8
= (1 − 4 sin x + 6 sin x − 4 sin x + sin x) cos x.

We rewrite the integral as

5 9 5 2 4 6 8
∫ sin x cos x dx = ∫ sin x(1 − 4 sin x + 6 sin x − 4 sin x + sin x) cos x dx.

Now substitute and integrate, using u = sin x and du = cos x dx .

5 2 4 6 8 5 7 9 11 13
∫ u (1 − 4 u + 6u − 4u + u ) du =∫ (u − 4u + 6u − 4u +u ) du

1 1 3 1 1
6 8 10 12 14
= u − u + u − u + u +C
6 2 5 3 14

1 6
1 8
3 10
= sin x− sin x+ sin x +…
6 2 5

1 12
1 14
− sin x+ sin x + C.
3 14

Technology Note: The work we are doing here can be a bit tedious, but the skills developed (problem solving, algebraic
manipulation, etc.) are important. Nowadays problems of this sort are often solved using a computer algebra system. The powerful
program Mathematica integrates ∫ sin x cos x dx as 5 9

45 cos(2x) 5 cos(4x) 19 cos(6x) cos(8x) cos(10x) cos(12x) cos(14x)


f (x) = − − + + − − − , (6.3.5)
16384 8192 49152 4096 81920 24576 114688

which clearly has a different form than our answer in Example 6.3.2, which is
1 6
1 8
3 10
1 12
1 14
g(x) = sin x− sin x+ sin x− sin x+ sin x. (6.3.6)
6 2 5 3 14

Figure 6.3.1 shows a graph of f and g ; they are clearly not equal, but they differ only by a constant. That is g(x) = f (x) + C for
some constant C . So we have two different antiderivatives of the same function, meaning both answers are correct.

6.3.2 https://math.libretexts.org/@go/page/4187
Figure 6.3.1 : A plot of f (x) and g(x) from Example 6.3.2 and the Technology Note.

Example 6.3.3: Integrating powers of sine and cosine

Evaluate ∫ cos 4 2
x sin x dx .
Solution
The powers of sine and cosine are both even, so we employ the power--reducing formulas and algebra as follows.
2
1 + cos(2x) 1 − cos(2x)
4 2
∫ cos x sin x dx = ∫ ( ) ( ) dx
2 2

2
1 + 2 cos(2x) + cos (2x) 1 − cos(2x)
=∫ ⋅ dx
4 2

1
2 3
=∫ (1 + cos(2x) − cos (2x) − cos (2x)) dx
8

The cos(2x) term is easy to integrate, especially with Key Idea 10. The cos (2x) term is another trigonometric integral with an
2

even power, requiring the power--reducing formula again. The cos (2x) term is a cosine function with an odd power, requiring
3

a substitution as done before. We integrate each in turn below.


1
∫ cos(2x) dx = sin(2x) + C .
2

1 + cos(4x) 1 1
2
∫ cos (2x) dx = ∫ dx = (x + sin(4x)) + C .
2 2 4

Finally, we rewrite cos 3


(2x) as
3 2 2
cos (2x) = cos (2x) cos(2x) = (1 − sin (2x)) cos(2x). (6.3.7)

Letting u = sin(2x), we have du = 2 cos(2x) dx, hence

3 2
∫ cos (2x) dx = ∫ (1 − sin (2x)) cos(2x) dx

1 2
=∫ (1 − u ) du
2

1 1
3
= (u − u )+C
2 3

1 1 3
= ( sin(2x) − sin (2x)) + C
2 3

Putting all the pieces together, we have

6.3.3 https://math.libretexts.org/@go/page/4187
4 2
1 2 3
∫ cos x sin x dx =∫ (1 + cos(2x) − cos (2x) − cos (2x)) dx
8

1 1 1 1 1 1
3
= [x + sin(2x) − (x + sin(4x)) − ( sin(2x) − sin (2x))] + C
8 2 2 4 2 3

1 1 1 1
3
= [ x− sin(4x) + sin (2x)] + C
8 2 8 6

The process above was a bit long and tedious, but being able to work a problem such as this from start to finish is important.

Integrals of the form ∫ sin(mx) sin(nx) dx, ∫ cos(mx) cos(nx) dx , and ∫ sin(mx) cos(nx) dx.

Functions that contain products of sines and cosines of differing periods are important in many applications including the analysis
of sound waves. Integrals of the form

∫ sin(mx) sin(nx) dx, ∫ cos(mx) cos(nx) dx and ∫ sin(mx) cos(nx) dx (6.3.8)

are best approached by first applying the Product to Sum Formulas found in the back cover of this text, namely
1
sin(mx) sin(nx) = [ cos ((m − n)x) − cos ((m + n)x)]
2

1
cos(mx) cos(nx) = [ cos ((m − n)x) + cos ((m + n)x)]
2

1
sin(mx) cos(nx) = [ sin ((m − n)x) + sin ((m + n)x)]
2

Example 6.3.4: Integrating products of sin(mx) and cos(nx)

Evaluate ∫ sin(5x) cos(2x) dx.


Solution
The application of the formula and subsequent integration are straightforward:
1
∫ sin(5x) cos(2x) dx = ∫ [ sin(3x) + sin(7x)] dx
2

1 1
=− cos(3x) − cos(7x) + C
6 14

Integrals of the form ∫ tan m


x sec
n
x dx .
When evaluating integrals of the form ∫ sin x cos x dx , the Pythagorean Theorem allowed us to convert even powers of sine
m n

into even powers of cosine, and vise--versa. If, for instance, the power of sine was odd, we pulled out one sin x and converted the
remaining even power of sin x into a function using powers of cos x, leading to an easy substitution.
The same basic strategy applies to integrals of the form ∫ tan
m
x sec
n
x dx , albeit a bit more nuanced. The following three facts
will prove useful:
1. (tan x) = sec x ,
d

dx
2

2. (sec x) = sec x tan x, and


d

dx

3. 1 + tan x = sec x (the Pythagorean Theorem).


2 2

If the integrand can be manipulated to separate a sec x term with the remaining secant power even, or if a sec x tan x term can be
2

separated with the remaining tan x power even, the Pythagorean Theorem can be employed, leading to a simple substitution. This
strategy is outlined in the following Key Idea.

6.3.4 https://math.libretexts.org/@go/page/4187
Key Idea 12: Integrals Involving Powers of Tangent and Secant
Consider ∫ tan m
x sec
n
x dx , where m, n are nonnegative integers.
1. If n is even, then n = 2k for some integer k . Rewrite sec n
x as
n 2k 2k−2 2 2 k−1 2
sec x = sec x = sec x sec x = (1 + tan x) sec x. (6.3.9)

Then

m n m 2 k−1 2 m 2 k−1
∫ tan x sec x dx = ∫ tan x(1 + tan x) sec x dx = ∫ u (1 + u ) du, (6.3.10)

where u = tan x and du = sec x dx . 2

2. If m is odd, then m = 2k + 1 for some integer k . Rewrite tan m


x sec
n
x as
m n 2k+1 n 2k n−1 2 k n−1
tan x sec x = tan x sec x = tan x sec x sec x tan x = (sec x − 1) sec x sec x tan x. (6.3.11)

Then

m n 2 k n−1 2 k n−1
∫ tan x sec x dx = ∫ (sec x − 1) sec x sec x tan x dx = ∫ (u − 1) u du, (6.3.12)

where u = sec x and du = sec x tan x dx.


3. If n is odd and m is even, then m = 2k for some integer k . Convert tan m
x to (sec 2
x − 1)
k
. Expand the new integrand
and use Integration By Parts, with dv = sec x dx . 2

4. If m is even and n = 0 , rewrite tan x as m

m m−2 2 m−2 2 m−2 2 m−2


tan x = tan x tan x = tan x(sec x − 1) = tan sec x − tan x. (6.3.13)

So

m m−2 2 m−2
∫ tan x dx = ∫ tan sec x dx − ∫ tan x dx . (6.3.14)

 
\small apply rule \#1 \small apply rule \#4 again

The techniques described in items 1 and 2 of Key Idea 12 are relatively straightforward, but the techniques in items 3 and 4 can be
rather tedious. A few examples will help with these methods.

Example 6.3.5: Integrating powers of tangent and secant

Evaluate ∫ tan 2
x sec
6
x dx .
Solution
Since the power of secant is even, we use rule #1 from Key Idea 12 and pull out a sec
2
x in the integrand. We convert the
remaining powers of secant into powers of tangent.

2 6 2 4 2
∫ tan x sec x dx = ∫ tan x sec x sec x dx

2 2 2 2
=∫ tan x(1 + tan x) sec x dx

Now substitute, with u = tan x , with du = sec 2


x dx .

2 2 2
=∫ u (1 + u ) du

We leave the integration and subsequent substitution to the reader. The final answer is
1 2 1
3 5 7
= tan x+ tan x+ tan x + C.
3 5 7

6.3.5 https://math.libretexts.org/@go/page/4187
Example 6.3.6: Integrating powers of tangent and secant

Evaluate ∫ sec 3
x dx .
Solution
We apply rule #3 from Key Idea 12 as the power of secant is odd and the power of tangent is even (0 is an even number). We
use Integration by Parts; the rule suggests letting dv = sec x dx , meaning that u = sec x .
2

u = sec x v =?

2
du = ? dv = sec x dx

Figure 6.3.2 : Setting up Integration by Parts.


Employing Integration by Parts, we have

3 2
∫ sec x dx = ∫ sec x ⋅ sec x dx
 
u dv

2
= sec x tan x − ∫ sec x tan x dx.

This new integral also requires applying rule \#3 of Key Idea:

2
= sec x tan x − ∫ sec x( sec x − 1) dx

3
= sec x tan x − ∫ sec x dx + ∫ sec x dx

3
= sec x tan x − ∫ sec x dx + ln | sec x + tan x|

In previous applications of Integration by Parts, we have seen where the original integral has reappeared in our work. We
resolve this by adding ∫ sec x dx to both sides, giving:
3

3
2∫ sec x dx = sec x tan x + ln | sec x + tan x|

3
1
∫ sec x dx = ( sec x tan x + ln | sec x + tan x|) + C
2

We give one more example.

Example 6.3.7: Integrating powers of tangent and secant

Evaluate ∫ tan 6
x dx .
Solution
We employ rule #4 of Key Idea 12.

6 4 2
∫ tan x dx = ∫ tan x tan x dx

4 2
=∫ tan x( sec x − 1) dx

4 2 4
=∫ tan x sec x dx − ∫ tan x dx

Integrate the first integral with substitution, u = tan x ; integrate the second by employing rule #4 again.

6.3.6 https://math.libretexts.org/@go/page/4187
1 5 2 2
= tan x −∫ tan x tan x dx
5

1 5 2 2
= tan x −∫ tan x( sec x − 1) dx
5

1 5 2 2 2
= tan x −∫ tan x sec x dx + ∫ tan x dx
5

Again, use substitution for the first integral and rule \#4 for the second.
1 5
1 3 2
= tan x− tan x +∫ ( sec x − 1) dx
5 3

1 1
5 3
= tan x− tan x + tan x − x + C .
5 3

These latter examples were admittedly long, with repeated applications of the same rule. Try to not be overwhelmed by the length
of the problem, but rather admire how robust this solution method is. A trigonometric function of a high power can be
systematically reduced to trigonometric functions of lower powers until all antiderivatives can be computed.
The next section introduces an integration technique known as Trigonometric Substitution, a clever combination of Substitution and
the Pythagorean Theorem.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of VMI
and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/
Integrated by Justin Marshall.

This page titled 6.3: Trigonometric Integrals is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory
Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

6.3.7 https://math.libretexts.org/@go/page/4187
6.4: Trigonometric Substitution
In Section 5.2 we defined the definite integral as the signed area under the curve." In that section we had not yet learned the
Fundamental Theorem of Calculus, so we evaluated special definite integrals which described nice, geometric shapes. For instance,
we were able to evaluate
3
− −−− − 9π
∫ √ 9 − x2 dx = (6.4.1)
−3
2

−−−−−
as we recognized that f (x) = √9 − x described the upper half of a circle with radius 3.
2

We have since learned a number of integration techniques, including Substitution and Integration by Parts, yet we are still unable to
evaluate the above integral without resorting to a geometric interpretation. This section introduces Trigonometric Substitution, a
method of integration that fills this gap in our integration skill. This technique works on the same principle as Substitution as found
in Section 6.1, though it can feel "backward." In Section 6.1, we set u = f (x), for some function f , and replaced f (x) with u. In
this section, we will set x = f (θ) , where f is a trigonometric function, then replace x with f (θ) .
We start by demonstrating this method in evaluating the integral in 6.4.1 . After the example, we will generalize the method and
give more examples.

Example 6.4.1: Using Trigonometric Substitution


3 −−−− −
Evaluate ∫ −3
√9 − x2 dx .
Solution
We begin by noting that 9 sin
2
θ + 9 cos
2
θ =9 , and hence 9 cos
2 2
θ = 9 − 9 sin θ . If we let x = 3 sin θ , then
9 −x
2
= 9 − 9 sin
2
θ = 9 cos
2
θ .
Setting x = 3 sin θ gives dx = 3 cos θ dθ . We are almost ready to substitute. We also wish to change our bounds of
integration. The bound x = −3 corresponds to θ = −π/2 (for when θ = −π/2 , x = 3 sin θ = −3 ). Likewise, the bound of
x = 3 is replaced by the bound θ = π/2 . Thus

3 π/2
− −−− − − −−− −− −−−
∫ √ 9 − x2 dx = ∫ √ 9 − 9 sin2 θ (3 cos θ) dθ (6.4.2)
−3 −π/2

π/2
−−−−−−
2
=∫ 3 √ 9 cos θ cos θ dθ (6.4.3)
−π/2

π/2

=∫ 3|3 cos θ| cos θ dθ. (6.4.4)


−π/2

On [−π/2, π/2], cos θ is always positive, so we can drop the absolute value bars, then employ a power--reducing formula:
π/2
2
=∫ 9 cos θ dθ (6.4.5)
−π/2

π/2
9
=∫ (1 + cos(2θ)) dθ (6.4.6)
−π/2
2

π/2
9 1 ∣ 9
= (θ + sin(2θ))∣ = π. (6.4.7)
2 2 ∣ 2
−π/2

This matches our answer from before.

−−−−−−
We now describe in detail Trigonometric Substitution. This method excels when dealing with integrands that contain √a − x , 2 2

−−−− −− −− −− −−
2
√x − a and √x + a . The following Key Idea 13 outlines the procedure for each case, followed by more examples. Each right
2 2 2

triangle acts as a reference to help us understand the relationships between x and θ .

6.4.1 https://math.libretexts.org/@go/page/4188
Key Idea 13: Trigonometric Substitution
−−−−−−
a. For integrands containing √a − x :
2 2

Let x = a sin θ , dx = a cos θ dθ


Thus θ = sin (x/a) , for −π/2 ≤ θ ≤ π/2 .
−1

On this interval, cos θ ≥ 0 , so


− − −− −−
= a cos θ .
2
√a − x 2

−−−−−−
b. For integrands containing √x + a :
2 2

Let x = a tan θ , dx = a sec θ dθ


2

Thus θ = tan (x/a) , for −π/2 < θ < π/2 .


−1

On this interval, sec θ > 0 , so


− − −− −−
= a sec θ .
2
√x + a 2

−−−−−−
c. For integrands containing √x − a :
2 2

Let x = a sec θ , dx = a sec θ tan θ dθ


Thus θ = sec (x/a) . If x/a ≥ 1 , then 0 ≤ θ < π/2 ; if x/a ≤ −1 , then π/2 < θ ≤ π .
−1

We restrict our work to where x ≥ a , so x/a ≥ 1 , and 0 ≤ θ < π/2 .


On this interval, tan θ ≥ 0 , so
− − −− −−
= a tan θ .
2
√x − a 2

Example 6.4.2: Using Trigonometric Substitution

Evaluate ∫ 1
dx.
√5+x2

Solution
– – –
Using Key Idea 13(b), we recognize a = √5 and set x = √5 tan θ . This makes dx = √5 sec 2
θ dθ . We will use the fact that
− −−−− − −−−−−−− − − −−− −− –
= √5 + 5 tan θ = √5 sec θ = √5 sec θ. Substituting, we have:
2 2 2
√5 + x

1 1 – 2
∫ − −−− − dx = ∫ − −−−−−−− − √5 sec θ dθ (6.4.8)
√ 5 + x2 2
√ 5 + 5 tan θ
– 2
√5 sec θ
=∫ – dθ (6.4.9)
√5 sec θ

=∫ sec θ dθ (6.4.10)

= ln ∣
∣ sec θ + tan θ∣
∣ + C. (6.4.11)

While the integration steps are over, we are not yet done. The original problem was stated in terms of x, whereas our answer is
given in terms of θ . We must convert back to x.

The reference triangle given in Key Idea 13(b) helps. With x = √5 tan θ , we have
− −−−−
x √ x2 + 5
tan θ = – and sec θ = – . (6.4.12)
√5 √5

This gives
1
∫ ∣ ∣
− −−− − dx = ln ∣ sec θ + tan θ∣ + C (6.4.13)
√ 5 + x2
− −−− −
∣ √ x2 + 5 x ∣
= ln∣ – + – ∣ + C. (6.4.14)
∣ √5 √5 ∣

We can leave this answer as is, or we can use a logarithmic identity to simplify it. Note:

6.4.2 https://math.libretexts.org/@go/page/4188
− −−−−
∣ √ x2 + 5 x ∣ ∣ 1 − −−−− ∣
2
ln∣ + ∣+C = ln∣ (√ x + 5 + x)∣ + C (6.4.15)
– – –
∣ √5 √5 ∣ ∣ √5 ∣

∣ 1 ∣ − −−− −
2
= ln∣ ∣√ x + 5 + x ∣
– ∣ + ln ∣ ∣+C (6.4.16)
∣ √5 ∣
− −−− −
2
∣√ x + 5 + x ∣
= ln ∣ ∣ + C, (6.4.17)


where the ln (1/ √5) term is absorbed into the constant C . (In Section 6.6 we will learn another way of approaching this
problem.)

Example 6.4.3: Using Trigonometric Substitution


−−−−−−
Evaluate ∫ √4x 2
− 1 dx .
Solution
−− −−−−
We start by rewriting the integrand so that it looks like √x 2
−a
2
for some value of a :
−−−−−−−−−−
− −−−−− 1
2 2
√ 4 x − 1 = √ 4 (x − ) (6.4.18)
4

−−−−−−−−−
2
1
2
= 2√ x −( ) (6.4.19)
2

So we have a = 1/2 , and following Key Idea 13(c), we set x = 1

2
sec θ , and hence dx = 1

2
sec θ tan θ dθ .
We now rewrite the integral with these substitutions:
−−−−−−−−−
2
− −−−−− 1
2 2
∫ √ 4 x − 1 dx = ∫ 2√ x −( ) dx (6.4.20)
2

−−−− −−−−−−
1 1 1
2
=∫ 2√ sec θ − ( sec θ tan θ) dθ (6.4.21)
4 4 2
−−−− −− −−−−−
1
2
=∫ √ (sec θ − 1) ( sec θ tan θ) dθ (6.4.22)
4
−−−−− −
1
2
=∫ √ tan θ( sec θ tan θ) dθ (6.4.23)
4

1
2
=∫ tan θ sec θ dθ (6.4.24)
2

1
2
= ∫ ( sec θ − 1) sec θ dθ (6.4.25)
2

1 3
= ∫ ( sec θ − sec θ) dθ. (6.4.26)
2

We integrated sec 3
θ in Example 6.3.6, finding its antiderivatives to be

3
1
∫ sec θ dθ = ( sec θ tan θ + ln | sec θ + tan θ|) + C . (6.4.27)
2

Thus
− −−−−− 1
2 3
∫ √ 4 x − 1 dx = ∫ ( sec θ − sec θ) dθ (6.4.28)
2

1 1
= ( ( sec θ tan θ + ln | sec θ + tan θ|) − ln | sec θ + tan θ|) + C (6.4.29)
2 2

1
= (sec θ tan θ − ln | sec θ + tan θ|) + C . (6.4.30)
4

6.4.3 https://math.libretexts.org/@go/page/4188
We are not yet done. Our original integral is given in terms of x, whereas our final answer, as given, is in terms of θ . We need
to rewrite our answer in terms of x. With a = 1/2 , and x = sec θ , the reference triangle in Key Idea 13(c) shows that
1

−−−−−−− −−−−−−−
2 2
tan θ = √ x − 1/4 /(1/2) = 2 √ x − 1/4 and sec θ = 2x. (6.4.31)

Thus
−−−−−−− −−−−−−−
1 1 2 2
( sec θ tan θ − ln ∣
∣ sec θ + tan θ∣
∣) + C = (2x ⋅ 2 √ x − 1/4 − ln ∣
∣2x + 2 √ x − 1/4 ∣
∣) + C (6.4.32)
4 4
−−−−−−− −−−−−−−
1 2 2
= (4x √ x − 1/4 − ln ∣
∣2x + 2 √ x − 1/4 ∣
∣) + C . (6.4.33)
4

The final answer is given in the last line above, repeated here:
− −−−−− 1 − −−−−−− − −−−−−−
2 2 2
∫ √ 4 x − 1 dx = (4x √ x − 1/4 − ln ∣
∣2x + 2 √ x − 1/4 ∣
∣) + C . (6.4.34)
4

Example 6.4.4: Using Trigonometric Substitution


√4−x2
Evaluate ∫ x2
dx .
Solution
−−−−−
We use Key Idea 13(a) with a = 2 , x = 2 sin θ , dx = 2 cos θ and hence √4 − x 2
= 2 cos θ . This gives
− −−− −
√ 4 − x2 2 cos θ
∫ dx = ∫ (2 cos θ) dθ (6.4.35)
2 2
x 4 sin θ

2
=∫ cot θ dθ (6.4.36)

2
= ∫ (csc θ − 1) dθ (6.4.37)

= − cot θ − θ + C . (6.4.38)

We need to rewrite our answer in terms of x . Using the reference triangle found in Key Idea 13(a), we have
−−−− −
cot θ = √4 − x /x and θ = sin (x/2) . Thus
2 −1

− −−− − − −−− −
√ 4 − x2 √ 4 − x2 x
−1
∫ dx = − − sin ( ) + C. (6.4.39)
2
x x 2

−−−− −− −−−−− − −−−−− −


Trigonometric Substitution can be applied in many situations, even those not of the form √a2 − x2 , √x2 − a2 or √x2 + a2. In
the following example, we apply it to an integral we already know how to handle.

Example 6.4.5: Using Trigonometric Substitution

Evaluate ∫ 1

x2 +1
dx .

Solution
We know the answer already as tan x + C . We apply Trigonometric Substitution here to show that we get the same answer
−1

without inherently relying on knowledge of the derivative of the arctangent function.


Using Key Idea 13(b), let x = tan θ , dx = sec 2
θ dθ and note that x 2
+ 1 = tan
2
θ + 1 = sec
2
θ . Thus
1 1 2
∫ dx = ∫ sec θ dθ (6.4.40)
2 2
x +1 sec θ

=∫ 1 dθ (6.4.41)

= θ + C. (6.4.42)

Since x = tan θ , θ = tan −1


x , and we conclude that ∫ 1

x2 +1
dx = tan
−1
x + C.

6.4.4 https://math.libretexts.org/@go/page/4188
The next example is similar to the previous one in that it does not involve a square--root. It shows how several techniques and
identities can be combined to obtain a solution.

Example 6.4.6: Using Trigonometric Substitution

Evaluate ∫ 1
2
dx.
( x2 +6x+10 )

Solution
We start by completing the square, then make the substitution u = x +3 , followed by the trigonometric substitution of
u = tan θ :

1 1 1
∫ dx = ∫ dx = ∫ du. (6.4.43)
2 2 2 2 2
(x + 6x + 10 ) 2 (u + 1)
((x + 3 ) + 1)

Now make the substitution u = tan θ , du = sec 2


θ dθ :
1 2
=∫ sec θ dθ (6.4.44)
2
(tan θ + 1 )2

1 2
=∫ sec θ dθ (6.4.45)
2 2
(sec θ)

2
=∫ cos θ dθ. (6.4.46)

Applying a power reducing formula, we have


1 1
=∫ ( + cos(2θ)) dθ (6.4.47)
2 2

1 1
= θ+ sin(2θ) + C . (6.4.48)
2 4

We need to return to the variable x. As u = tan θ , θ = tan


−1
u . Using the identity sin(2θ) = 2 sin θ cos θ and using the
reference triangle found in Key Idea 13(b), we have
1 1 u 1 1 u
sin(2θ) = − −−−− ⋅ −−−−− = . (6.4.49)
2
4 2 √ u2 + 1 2
√u + 1 2 u +1

Finally, we return to x with the substitution u = x + 3 . We start with the expression in Equation (???) :
1 1 1 1 u
−1
θ+ sin(2θ) + C = tan u+ +C (6.4.50)
2
2 4 2 2 u +1

1 x +3
−1
= tan (x + 3) + + C. (6.4.51)
2
2 2(x + 6x + 10)

Stating our final result in one line,


1 1 −1
x +3
∫ dx = tan (x + 3) + + C. (6.4.52)
(x2 + 6x + 10 )2 2 2(x2 + 6x + 10)

Our last example returns us to definite integrals, as seen in our first example. Given a definite integral that can be evaluated using
Trigonometric Substitution, we could first evaluate the corresponding indefinite integral (by changing from an integral in terms of
x to one in terms of θ , then converting back to x) and then evaluate using the original bounds. It is much more straightforward,

though, to change the bounds as we substitute.

Example 6.4.7: Definite integration and Trigonometric Substitution


2
5
Evaluate ∫ 0
x
dx .
√x2 +25

Solution

6.4.5 https://math.libretexts.org/@go/page/4188
−−−−− −
Using Key Idea 13(b), we set x = 5 tan θ , dx = 5 sec 2
θ dθ , and note that √x 2
+ 25 = 5 sec θ . As we substitute, we can also
change the bounds of integration.
The lower bound of the original integral is x = 0 . As x = 5 tan θ , we solve for θ and find θ = tan (x/5) . Thus the new −1

lower bound is θ = tan (0) = 0 . The original upper bound is x = 5 , thus the new upper bound is θ = tan (5/5) = π/4 .
−1 −1

Thus we have
5 2 π/4 2
x 25 tan θ
2
∫ − −−−− − dx = ∫ 5 sec θ dθ (6.4.53)
0 √ x2 + 25 0
5 sec θ

π/4
2
= 25 ∫ tan θ sec θ dθ. (6.4.54)
0

We encountered this indefinite integral in Example 6.4.3 where we found

2
1
∫ tan θ sec θ dθ = ( sec θ tan θ − ln | sec θ + tan θ|). (6.4.55)
2

So
π/4 π/4
25 ∣
2
25 ∫ tan θ sec θ dθ = ( sec θ tan θ − ln | sec θ + tan θ|)∣ (6.4.56)
0
2 ∣
0

25 – –
= (√2 − ln(√2 + 1)) (6.4.57)
2

≈ 6.661. (6.4.58)

The following equalities are very useful when evaluating integrals using Trigonometric Substitution.

Ket Idea 14: Useful Equalities with Trigonometric Substitution


1. sin(2θ) = 2 sin θ cos θ
2. cos(2θ) = cos θ − sin2 2
θ = 2 cos
2
θ − 1 = 1 − 2 sin
2
θ

3. ∫ sec 3
θ dθ =
1

2
( sec θ tan θ + ln ∣
∣ sec θ + tan θ∣
∣) + C

4. ∫ cos 2
θ dθ = ∫
1

2
(1 + cos(2θ)) dθ =
1

2
(θ + sin θ cos θ) + C .

The next section introduces Partial Fraction Decomposition, which is an algebraic technique that turns "complicated" fractions into
sums of "simpler" fractions, making integration easier.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/
Integrated by Justin Marshall.

This page titled 6.4: Trigonometric Substitution is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory
Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

6.4.6 https://math.libretexts.org/@go/page/4188
6.5: Partial Fraction Decomposition
In this section we investigate the antiderivatives of rational functions. Recall that rational functions are functions of the form
p(x)
f (x) =
q(x)
, where p(x) and q(x) are polynomials and q(x) ≠ 0 . Such functions arise in many contexts, one of which is the
solving of certain fundamental differential equations.
We begin with an example that demonstrates the motivation behind this section. Consider the integral ∫ 1
2
x −1
dx . We do not have a
simple formula for this (if the denominator were x + 1 , we would recognize the antiderivative as being the arctangent function). It
2

can be solved using Trigonometric Substitution, but note how the integral is easy to evaluate once we realize:
$$\frac{1}{x^2-1} = \frac{1/2}{x-1} - \frac{1/2}{x+1}.\]
Thus

1 1/2 1/2
∫ dx = ∫ dx − ∫ dx (6.5.1)
x2 − 1 x −1 x +1

1 1
= ln |x − 1| − ln |x + 1| + C . (6.5.2)
2 2

This section teaches how to decompose


$$\frac{1}{x^2-1}\quad \text{into}\quad \frac{1/2}{x-1}-\frac{1/2}{x+1}.\]
p(x)
We start with a rational function f (x) = q(x)
, where p and q do not have any common factors and the degree of p is less than the
degree of q. It can be shown that any polynomial, and hence q, can be factored into a product of linear and irreducible quadratic
terms. The following Key Idea states how to decompose a rational function into a sum of rational functions whose denominators
are all of lower degree than q.

Key Idea 15: Partial Fraction Decomposition


p(x)
Let q(x)
be a rational function, where the degree of p is less than the degree of q.

1. Linear Terms: Let (x − a) divide q(x), where (x − a) is the highest power of (x − a) that divides q(x). Then the
n

p(x)
decomposition of q(x)
will contain the sum

A1 A2 An
+ +⋯ + . (6.5.3)
2 n
(x − a) (x − a) (x − a)

2. Quadratic Terms: Let x 2


+ bx + c divide q(x), where (x 2
+ bx + c )
n
is the highest power of x 2
+ bx + c that divides
p(x)
q(x) . Then the decomposition of q(x)
will contain the sum

B1 x + C1 B2 x + C2 Bn x + Cn
+ +⋯ + . (6.5.4)
2 2 2 2 n
x + bx + c (x + bx + c ) (x + bx + c )

To find the coefficients A , B and C :


i i i

1. Multiply all fractions by q(x), clearing the denominators. Collect like terms.
2. Equate the resulting coefficients of the powers of x and solve the resulting system of linear equations.

The following examples will demonstrate how to put this Key Idea into practice. Example 6.5.1 stresses the decomposition aspect
of the Key Idea.

Example 6.5.1: Decomposing into partial fractions

Decompose f (x) = 3 2
1

2 2
without solving for the resulting coefficients.
(x+5)(x−2 ) ( x +x+2)( x +x+7 )

Solution
The denominator is already factored, as both x + x + 2 and x + x + 7 cannot be factored further. We need to decompose
2 2

f (x) properly. Since (x + 5) is a linear term that divides the denominator, there will be a

6.5.1 https://math.libretexts.org/@go/page/4189
$$\frac{A}{x+5}\]
term in the decomposition.
As (x − 2) divides the denominator, we will have the following terms in the decomposition:
3

$$\frac{B}{x-2},\quad \frac{C}{(x-2)^2}\quad \text{and}\quad \frac{D}{(x-2)^3}.\]


The x 2
+x +2 term in the denominator results in a Ex+F
2
x +x+2
term.

Finally, the (x 2
+ x + 7)
2
term results in the terms
$$\frac{Gx+H}{x^2+x+7}\quad \text{and}\quad \frac{Ix+J}{(x^2+x+7)^2}.\]
All together, we have
1 A B C D
= + + + + (6.5.5)
3 2 2 2 2 3
(x + 5)(x − 2 ) (x + x + 2)(x + x + 7) x +5 x −2 (x − 2) (x − 2)

Ex + F Gx + H Ix + J
+ + (6.5.6)
2 2 2 2
x +x +2 x +x +7 (x + x + 7)

Solving for the coefficients A , B … J would be a bit tedious but not "hard."

Example 6.5.2: Decomposing into partial fractions

Perform the partial fraction decomposition of 1


2
x −1
.

Solution
The denominator factors into two linear terms: x 2
− 1 = (x − 1)(x + 1) . Thus
$$\frac{1}{x^2-1} = \frac{A}{x-1} + \frac{B}{x+1}.\]
To solve for A and B , first multiply through by x 2
− 1 = (x − 1)(x + 1) :
A(x − 1)(x + 1) B(x − 1)(x + 1)
1 = + (6.5.7)
x −1 x +1

= A(x + 1) + B(x − 1) (6.5.8)

= Ax + A + Bx − B (6.5.9)

Now collect like terms.


= (A + B)x + (A − B). (6.5.10)

The next step is key. Note the equality we have:


1 = (A + B)x + (A − B). (6.5.11)

For clarity's sake, rewrite the left hand side as


0x + 1 = (A + B)x + (A − B). (6.5.12)

On the left, the coefficient of the x term is 0; on the right, it is (A + B) . Since both sides are equal, we must have that
0 = A+B .

Likewise, on the left, we have a constant term of 1; on the right, the constant term is (A − B) . Therefore we have 1 = A − B .
We have two linear equations with two unknowns. This one is easy to solve by hand, leading to

A+B = 0 A = 1/2
⇒ . (6.5.13)
A−B = 1 B = −1/2

Thus $$\frac{1}{x^2-1} = \frac{1/2}{x-1}-\frac{1/2}{x+1}.\]

6.5.2 https://math.libretexts.org/@go/page/4189
Example 6.5.3: Integrating using partial fractions

Use partial fraction decomposition to integrate ∫ 1


2
dx.
(x−1)(x+2)

Solution
We decompose the integrand as follows, as described by Key Idea 15:
$$\frac{1}{(x-1)(x+2)^2} = \frac{A}{x-1} + \frac{B}{x+2} + \frac{C}{(x+2)^2}.\]
To solve for A , B and C , we multiply both sides by (x − 1)(x + 2) and collect like terms:
2

2
1 = A(x + 2 ) + B(x − 1)(x + 2) + C (x − 1) (6.5.14)
2 2
= Ax + 4Ax + 4A + Bx + Bx − 2B + C x − C (6.5.15)
2
= (A + B)x + (4A + B + C )x + (4A − 2B − C ) (6.5.16)

Note: Equation 6.5.22 offers a direct route to finding the values of A , B and C . Since the equation holds for all values of x, it
holds in particular when x = 1 . However, when x = 1 , the right hand side simplifies to A(1 + 2) = 9A . Since the left hand
2

side is still 1, we have 1 = 9A . Hence A = 1/9 .


Likewise, the equality holds when x = −2 ; this leads to the equation 1 = −3C . Thus C = −1/3 .
Knowing A and C , we can find the value of B by choosing yet another value of x, such as x = 0 , and solving for B .
We have
$$0x^2+0x+ 1 = (A+B)x^2 + (4A+B+C)x + (4A-2B-C)\]
leading to the equations
$$A+B = 0, \quad 4A+B+C = 0 \quad \text{and} \quad 4A-2B-C = 1.\]
These three equations of three unknowns lead to a unique solution:
$$A = 1/9,\quad B = -1/9 \quad \text{and} \quad C = -1/3.\]
Thus
$$\int\frac{1}{(x-1)(x+2)^2}\ dx = \int \frac{1/9}{x-1}\ dx + \int \frac{-1/9}{x+2}\ dx + \int \frac{-1/3}{(x+2)^2}\ dx.\]
Each can be integrated with a simple substitution with u = x −1 or u = x +2 (or by directly applying Key Idea 10 as the
denominators are linear functions). The end result is
$$\int\frac{1}{(x-1)(x+2)^2}\ dx = \frac19\ln|x-1| -\frac19\ln|x+2| +\frac1{3(x+2)}+C.\]

Example 6.5.4: Integrating using partial fractions


3

Use partial fraction decomposition to integrate ∫ x

(x−5)(x+3)
dx .

Solution
Key Idea 15 presumes that the degree of the numerator is less than the degree of the denominator. Since this is not the case
here, we begin by using polynomial division to reduce the degree of the numerator. We omit the steps, but encourage the reader
to verify that
$$\frac{x^3}{(x-5)(x+3)} = x+2+\frac{19x+30}{(x-5)(x+3)}.\]
Using Key Idea 15, we can rewrite the new rational function as:
$$\frac{19x+30}{(x-5)(x+3)} = \frac{A}{x-5} + \frac{B}{x+3}\]
for appropriate values of A and B . Clearing denominators, we have
Note: The values of A and B can be quickly found using the technique described in the margin of Example 6.5.3.}
19x + 30 = A(x + 3) + B(x − 5) (6.5.17)

= (A + B)x + (3A − 5B). (6.5.18)

6.5.3 https://math.libretexts.org/@go/page/4189
This implies that:
19 = A + B (6.5.19)

30 = 3A − 5B. (6.5.20)

Solving this system of linear equations gives


125/8 = A (6.5.21)

27/8 = B. (6.5.22)

We can now integrate.


3
x 125/8 27/8
∫ dx =∫ (x + 2 + + ) dx (6.5.23)
(x − 5)(x + 3) x −5 x +3

2
x 125 27
= + 2x + ln |x − 5| + ln |x + 3| + C . (6.5.24)
2 8 8

Example 6.5.5: Integrating using partial fractions


2

Use partial fraction decomposition to evaluate ∫ 7 x +31x+54


2
(x+1)( x +6x+11)
dx.

Solution
The degree of the numerator is less than the degree of the denominator so we begin by applying Key Idea 15. We have:
2
7x + 31x + 54 A Bx + C
= + . (6.5.25)
2 2
(x + 1)(x + 6x + 11) x +1 x + 6x + 11

Now clear the denominators.


2 2
7x + 31x + 54 = A(x + 6x + 11) + (Bx + C )(x + 1) (6.5.26)

2
= (A + B)x + (6A + B + C )x + (11A + C ). (6.5.27)

This implies that:


7 = A+B (6.5.28)

31 = 6A + B + C (6.5.29)

54 = 11A + C . (6.5.30)

Solving this system of linear equations gives the nice result of A = 5 , B = 2 and C = −1 . Thus
$$\int\frac{7x^2+31x+54}{(x+1)(x^2+6x+11)}\ dx = \int\left(\frac{5}{x+1} + \frac{2x-1}{x^2+6x+11}\right)\ dx.\]
The first term of this new integrand is easy to evaluate; it leads to a 5 ln |x + 1| term. The second term is not hard, but takes
several steps and uses substitution techniques.
The integrand 2
2x−1

x +6x+11
has a quadratic in the denominator and a linear term in the numerator. This leads us to try substitution.
Let u = x + 6x + 11 , so du = (2x + 6) dx . The numerator is
2
2x − 1 , not 2x + 6 , but we can get a 2x + 6 term in the
numerator by adding 0 in the form of "7 − 7 ."
2x − 1 2x − 1 + 7 − 7
= (6.5.31)
2 2
x + 6x + 11 x + 6x + 11

2x + 6 7
= − . (6.5.32)
2 2
x + 6x + 11 x + 6x + 11

We can now integrate the first term with substitution, leading to a ln | x


2
+ 6x + 11| term. The final term can be integrated
using arctangent. First, complete the square in the denominator:
$$\frac{7}{x^2+6x+11} = \frac{7}{(x+3)^2+2}.\]
An antiderivative of the latter term can be found using Theorem 6.1.3 and substitution:

6.5.4 https://math.libretexts.org/@go/page/4189
$$\int \frac{7}{x^2+6x+11}\ dx = \frac{7}{\sqrt{2}}\tan^{-1}\left(\frac{x+3}{\sqrt{2}}\right)+C.\]
Let's start at the beginning and put all of the steps together.
2
7x + 31x + 54 5 2x − 1
∫ dx = ∫ ( + ) dx (6.5.33)
2 2
(x + 1)(x + 6x + 11) x +1 x + 6x + 11

5 2x + 6 7
=∫ dx + ∫ dx − ∫ dx (6.5.34)
2 2
x +1 x + 6x + 11 x + 6x + 11

7 x +3
2 −1
= 5 ln |x + 1| + ln | x + 6x + 11| − tan ( ) + C. (6.5.35)
– –
√2 √2

As with many other problems in calculus, it is important to remember that one is not expected to "see" the final answer
immediately after seeing the problem. Rather, given the initial problem, we break it down into smaller problems that are easier
to solve. The final answer is a combination of the answers of the smaller problems.

Partial Fraction Decomposition is an important tool when dealing with rational functions. Note that at its heart, it is a technique of
algebra, not calculus, as we are rewriting a fraction in a new form. Regardless, it is very useful in the realm of calculus as it lets us
evaluate a certain set of "complicated" integrals.
The next section introduces new functions, called the Hyperbolic Functions. They will allow us to make substitutions similar to
those found when studying Trigonometric Substitution, allowing us to approach even more integration problems.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/
Integrated by Justin Marshall.

This page titled 6.5: Partial Fraction Decomposition is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by
Gregory Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available
upon request.

6.5.5 https://math.libretexts.org/@go/page/4189
6.6: Hyperbolic Functions
The hyperbolic functions are a set of functions that have many applications to mathematics, physics, and engineering. Among
many other applications, they are used to describe the formation of satellite rings around planets, to describe the shape of a rope
hanging from two points, and have application to the theory of special relativity. This section defines the hyperbolic functions and
describes many of their properties, especially their usefulness to calculus.
These functions are sometimes referred to as the "hyperbolic trigonometric functions" as there are many, many connections
between them and the standard trigonometric functions. Figure 6.6.1 demonstrates one such connection. Just as cosine and sine are
used to define points on the circle defined by x + y = 1 , the functions hyperbolic cosine and hyperbolic sine are used to define
2 2

points on the hyperbola x − y = 1 .2 2

Figure 6.6.1 : Using trigonometric functions to define points on a circle and hyperbolic functions to define points on a hyperbola.
The area of the shaded regions are included in them.
We begin with their definition.

Definition 6.6.1: Hyperbolic Functions


x −x

1. cosh x = e +e

2
x −x

2. sinh x = e −e

3. tanh x = sinh x

cosh x

4. sechx = 1

cosh x

5. cschx = 1

sinh x

6. coth x = cosh x

sinh x

These hyperbolic functions are graphed in Figure 6.6.2. In the graphs of cosh x and sinh x, graphs of e /2 and e
x −x
are included
/2

with dashed lines. As x gets "large," cosh x and sinh x each act like e /2; when x is a large negative number,
x
cosh x acts like

/2 whereas $\sinh x$ acts like −e /2.


−x −x
e

Notice the domains of tanh x and sechx are (−∞, ∞), whereas both coth x and cschx have vertical asymptotes at x = 0 . Also
note the ranges of these functions, especially tanh x: as x → ∞ , both sinh x and cosh x approach e /2, hence tanh x
−x

approaches 1.
The following example explores some of the properties of these functions that bear remarkable resemblance to the properties of
their trigonometric counterparts.
Pronunciation Note: "cosh" rhymes with "gosh," "sinh" rhymes with "pinch," and "tanh" rhymes with "ranch."

6.6.1 https://math.libretexts.org/@go/page/4190
Figure 6.6.2 : Graphs of the hyperbolic functions.

Example 6.6.1: Exploring properties of hyperbolic functions

Use Definition 6.6.1 to rewrite the following expressions.


1. cosh x − sinh
2 2
x

2. tanh x + sech
2 2
x

3. 2 cosh x sinh x
4. ( cosh x)
d

dx

5. ( sinh x)
d

dx

6. ( tanh x)
d

dx

Solution

1. 2 2
e
x
+e
−x 2
e
x
−e
−x 2

cosh x − sinh x =( ) −( ) (6.6.1)


2 2

2x x −x −2x 2x x −x −2x
e + 2e e +e e − 2e e +e
= − (6.6.2)
4 4
4
= = 1. (6.6.3)
4

So cosh 2
x − sinh
2
x =1 .

2. 2 2
sinh
2
x 1
tanh x + sech x = + (6.6.4)
2 2
cosh x cosh x
2
sinh x +1
= Now use identity from #1. (6.6.5)
2
cosh x
2
cosh x
= = 1. (6.6.6)
2
cosh x

So tanh 2 2
x + sech x = 1 .

6.6.2 https://math.libretexts.org/@go/page/4190
3. e
x
+e
−x
e
x
−e
−x

2 cosh x sinh x = 2 ( )( ) (6.6.7)


2 2

2x −2x
e −e
=2⋅ (6.6.8)
4
2x −2x
e −e
= = sinh(2x). (6.6.9)
2

Thus 2 cosh x sinh x = sinh(2x).

4. d d e
x
+e
−x

( cosh x) = ( ) (6.6.10)
dx dx 2
x −x
e −e
= (6.6.11)
2

= sinh x. (6.6.12)

So d

dx
( cosh x) = sinh x.

5. d d e
x
−e
−x

( sinh x) = ( ) (6.6.13)
dx dx 2
x −x
e +e
= (6.6.14)
2

= cosh x. (6.6.15)

So d

dx
( sinh x) = cosh x.

6. d d sinh x
( tanh x) = ( ) (6.6.16)
dx dx cosh x

cosh x cosh x − sinh x sinh x


= (6.6.17)
2
cosh x
1
= (6.6.18)
2
cosh x
2
= sech x. (6.6.19)

So d

dx
( tanh x) = sech x.
2

The following Key Idea summarizes many of the important identities relating to hyperbolic functions. Each can be verified by
referring back to Definition 6.6.1.

Key Idea 16: Useful Hyperbolic Function Properties


Basic Identities
1. cosh x − sinh x = 1
2 2

2. tanh x + sech x = 1
2 2

3. coth x − csch x = 1
2 2

4. cosh 2x = cosh x + sinh 2 2


x

5. sinh 2x = 2 sinh x cosh x


6. cosh x =2 cosh 2x+1

2
cosh 2x−1
7. sinh 2
x =
2

Derivatives
1. d

dx
( cosh x) = sinh x

2. d

dx
( sinh x) = cosh x

3. d

dx
( tanh x) = sech x
2

4. d

dx
(sechx) = −sechx tanh x

5. d

dx
(cschx) = −cschx coth x

6.6.3 https://math.libretexts.org/@go/page/4190
6. d

dx
2
( coth x) = −csch x

Integrals
1. ∫ cosh x dx = sinh x + C
2. ∫ sinh x dx = cosh x + C
3. ∫ tanh x dx = ln(cosh x) + C
4. ∫ coth x dx = ln | sinh x | + C

We practice using Key Idea 16

Example 6.6.2: Derivatives and integrals of hyperbolic functions

Evaluate the following derivatives and integrals.


1. ( cosh 2x)
d

dx

2. ∫ sech (7t − 3) dt
2

ln 2
3. ∫ 0
cosh x dx

Solution
1. Using the Chain Rule directly, we have ( cosh 2x) = 2 sinh 2x.
d

dx

Just to demonstrate that it works, let's also use the Basic Identity found in Key Idea 16: cosh 2x = cosh 2
x + sinh
2
x .
d d
2 2
( cosh 2x) = ( cosh x + sinh x) = 2 cosh x sinh x + 2 sinh x cosh x (6.6.20)
dx dx

= 4 cosh x sinh x. (6.6.21)

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 10 and 16 we have:

2
1
∫ sech (7t − 3) dt = tanh(7t − 3) + C . (6.6.22)
7

3. ln 2


ln 2

∫ cosh x dx = sinh x = sinh(ln 2) − sinh 0 = sinh(ln 2). (6.6.23)



0
0

We can simplify this last expression as sinh x is based on exponentials:


ln 2 − ln 2
e −e 2 − 1/2 3
sinh(ln 2) = = = . (6.6.24)
2 2 4

Inverse Hyperbolic Functions


Just as the inverse trigonometric functions are useful in certain integrations, the inverse hyperbolic functions are useful with others.
Figure 16 shows the restrictions on the domains to make each function one-to-one and the resulting domains and ranges of their
inverse functions. Their graphs are shown in Figure 6.6.3
Because the hyperbolic functions are defined in terms of exponential functions, their inverses can be expressed in terms of
− −−− −
logarithms as shown in Key Idea 17. It is often more convenient to refer to sinh x than to ln (x + √x + 1 ) , especially when
−1 2

one is working on theory and does not need to compute actual values. On the other hand, when computations are needed,
technology is often helpful but many hand-held calculators lack a \textit{convenient} sinh x button. (Often it can be accessed −1

under a menu system, but not conveniently.) In such a situation, the logarithmic representation is useful. The reader is not
encouraged to memorize these, but rather know they exist and know how to use them when needed.

6.6.4 https://math.libretexts.org/@go/page/4190
Table 6.6.1 : Graphs of cosh x, sinh x and their inverses.

Figure 6.6.3 : Graphs of the hyperbolic functions and their inverses.


The following Key Ideas give the derivatives and integrals relating to the inverse hyperbolic functions. In Key Idea 19, both the
inverse hyperbolic and logarithmic function representations of the antiderivative are given, based on Key Idea 17. 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 6.4.

Key IDea 17: Logarithmic definitions of the inverse hyperbolic functions.


−−−−−
1. cosh −1
x = ln (x + √x2 − 1 ); x ≥ 1

2. tanh −1
x =
1

2
ln(
1+x

1−x
); |x| < 1

2
1+√1−x
3. sech −1
x = ln(
x
); 0 < x ≤ 1

−−−−−
4. sinh −1 2
x = ln (x + √x + 1 )

5. coth −1
x =
1

2
ln(
x+1

x−1
); |x| > 1

√1+x2
6. csch −1
x = ln(
1

x
+
|x|
); x ≠ 0

Key Idea 18: Derivatives Involving Inverse Hyperbolic Functions


1. d

dx
( cosh
−1
x) =
1
; x >1
√x2 −1

2. d

dx
( sinh
−1
x) =
1

√x2 +1

3. d

dx
( tanh
−1
x) =
1−x
1
2
; |x| < 1

−1
4. d

dx
(sech
−1
x) = ;0 <x <1
x√1−x2

6.6.5 https://math.libretexts.org/@go/page/4190
5. d

dx
(csch
−1
x) =
−1

2
; x ≠0
|x| √1+x

6. d

dx
( coth
−1
x) =
1

1−x
2
; |x| > 1

Key Idea 19: Integrals Involving Inverse Hyperbolic Functions


−− −−−−
1. ∫ 1
dx = cosh
−1
(
x

a
) + C; 0 < a < x


2
= ln x + √x − a
2 ∣

+C
√x2 −a2

−− −−−−
2. ∫ 1
dx = sinh
−1
(
x

a
) + C; a > 0


2
= ln x + √x + a
2 ∣

+C
√x2 +a2

1 −1 x 2 2

⎪ tanh ( )+C x <a
⎪ a a

3. ∫ 1

a2 −x2
dx = ⎨ =
2a
1
ln∣

a+x

a−x
∣+C



⎪ 1 −1 x 2 2
coth ( )+C a <x
a a

4. ∫ 2 2
1
dx = −
1

a
sech
−1
(
x

a
) + C; 0 < x < a =
1

a
ln(
2
x

2
) +C
x√a −x a+√a −x

∣ ∣
5. ∫ 1
dx = −
1

a
csch
−1


x

a

∣ + C ; x ≠ 0, a > 0 =
1

a
ln∣
x
∣+C
2
x√x +a
2
∣ 2
a+√a +x
2

We practice using the derivative and integral formulas in the following example.

Example 6.6.3: Derivatives and integrals involving inverse hyperbolic functions

Evaluate the following.

1. d

dx
[cosh
−1
(
3x−2

5
)]

2. ∫ 2
1

x −1
dx

3. ∫ 1
dx
√9 x2 +10

Solution
1. Applying Key Idea 18 with the Chain Rule gives:
d 3x − 2 1 3
−1
[cosh ( )] = ⋅ . (6.6.25)
−−−−−−−−−−
dx 5 2 5
3x−2
√( ) −1
5

−1
2. Multiplying the numerator and denominator by (−1) gives: ∫ dx = ∫
x −1
2
1

1−x
2
dx . The second integral can be solved
with a direct application of item #3 from Key Idea 19, with a = 1 . Thus
1 1
∫ dx = − ∫ dx (6.6.26)
2 2
x −1 1 −x
−1 2
⎧ − tanh (x) + C x <1

=⎨ (6.6.27)

⎪ −1 2
− coth (x) + C 1 <x

1
∣ x +1 ∣
=− ln∣ ∣+C (6.6.28)
2 ∣ x −1 ∣

1 ∣ x −1 ∣
= ln∣ ∣ + C. (6.6.29)
2 ∣ x +1 ∣

We should note that this exact problem was solved at the beginning of Section 6.5. In that example the answer was given as
1

2
ln |x − 1| −
1
ln |x + 1| + C .
2
Note that this is equivalent to the answer given in Equation 6.6.29, as
ln(a/b) = ln a − ln b .

3. This requires a substitution, then item #2 of Key Idea 19 can be applied.


Let u = 3x, hence du = 3dx . We have
1 1 1
∫ − − −−−− − dx = ∫ − −−−− − du. (6.6.30)
√ 9 x2 + 10 3 √ u2 + 10

6.6.6 https://math.libretexts.org/@go/page/4190
−−
Note a 2
= 10 , hence a = √10. Now apply the integral rule.
1 −1
3x
= sinh ( −−) +C (6.6.31)
3 √10

1 − − −−−−−
∣ 2 ∣
= ln 3x + √ 9 x + 10 + C . (6.6.32)
∣ ∣
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 memorized, but rather as a reference for future problem
solving. Key Idea 19 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 6.8, where limits will arise in the
evaluation of certain definite integrals.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/
Integrated by Justin Marshall.

This page titled 6.6: Hyperbolic Functions is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory
Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

6.6.7 https://math.libretexts.org/@go/page/4190
6.7: L'Hopital's Rule
While this chapter is devoted to learning techniques of integration, this section is not about integration. Rather, it is concerned with
a technique of evaluating certain limits that will be useful in the following section, where integration is once more discussed.
Our treatment of limits exposed us to "0/0", an indeterminate form. If lim f (x) = 0 and lim g(x) = 0 , we do not conclude
x→c x→c

that lim x→cf (x)/g(x) is 0/0; rather, we use 0/0 as notation to describe the fact that both the numerator and denominator

approach 0. The expression 0/0 has no numeric value; other work must be done to evaluate the limit.
Other indeterminate forms exist; they are: ∞/∞, 0 ⋅ ∞ , ∞ − ∞ , 0 , 1 and ∞ . Just as "0/0" does not mean "divide 0 by 0," the
0 ∞ 0

expression "∞/∞" does not mean "divide infinity by infinity." Instead, it means "a quantity is growing without bound and is being
divided by another quantity that is growing without bound." We cannot determine from such a statement what value, if any, results
in the limit. Likewise, "0 ⋅ ∞ " does not mean "multiply zero by infinity." Instead, it means "one quantity is shrinking to zero, and is
being multiplied by a quantity that is growing without bound." We cannot determine from such a description what the result of such
a limit will be.
This section introduces L'Hôpital's Rule, a method of resolving limits that produce the indeterminate forms 0/0 and ∞/∞. We'll
also show how algebraic manipulation can be used to convert other indeterminate expressions into one of these two forms so that
our new rule can be applied.

Theorem 6.7.1: L'Hôpital's Rule

Let lim x→c f (x) = 0 and lim g(x) = 0 , where f and g are differentiable functions on an open interval I containing c , and
x→c


g (x) ≠ 0 on I except possibly at c . Then
$$ \lim_{x\to c} \frac{f(x)}{g(x)} = \lim_{x\to c} \frac{f'(x)}{g'(x)}.\]

We demonstrate the use of L'Hôpital's Rule in the following examples; we will often use "LHR" as an abbreviation of "L'Hôpital's"

Example 6.7.1: Using L'Hôpital's Rule

Evaluate the following limits, using L'Hôpital's Rule as needed.


1. lim x→0
sin x

x
√x+3 −2
2. lim x→1
1−x
2

3. lim x→0
x

1−cos x
2

4. lim x→2
x +x−6
2
x −3x+2

Solution
1. We proved this limit is 1 in Example ??? using the Squeeze Theorem. Here we use L'Hôpital's Rule to show its power.
sin x by LHR cos x
lim = lim = 1. (6.7.1)
x→0 x x→0 1

1 −1/2
(x+3 )
√x+3 −2 by LHR
2. lim x→1
1−x
= limx→1
2

−1
=−
1

4
.

2 by LHR
3. lim x→0
x

1−cos x
= lim .
x→0
2x

sin x

This latter limit also evaluates to the 0/0 indeterminate form. To evaluate it, we apply L'Hôpital's Rule again.
2x by LHR 2
lim = = 2. (6.7.2)
x→0 sin x cos x

Thus lim x→0 = 2.


x

1−cos x

4. We already know how to evaluate this limit; first factor the numerator and denominator. We then have:
2
x +x −6 (x − 2)(x + 3) x +3
lim = lim = lim = 5. (6.7.3)
2
x→2 x − 3x + 2 x→2 (x − 2)(x − 1) x→2 x −1

We now show how to solve this using L'Hôpital's Rule.

6.7.1 https://math.libretexts.org/@go/page/4659
2
x +x −6 by LHR 2x + 1
lim = lim = 5. (6.7.4)
2
x→2 x − 3x + 2 x→2 2x − 3

Note that at each step where L'Hôpital's Rule was applied, it was needed: the initial limit returned the indeterminate form of "0/0."
If the initial limit returns, for example, 1/2, then L'Hôpital's Rule does not apply.
The following theorem extends our initial version of L'Hôpital's Rule in two ways. It allows the technique to be applied to the
indeterminate form ∞/∞ and to limits where x approaches ±∞ .

Theorem 6.7.2: L'Hôpital's Rule, Part 2

1. Let lim x→a f (x) = ±∞ and lim x→a g(x) = ±∞ , where f and g are differentiable on an open interval I containing a .
Then

f (x) f (x)
lim = lim . (6.7.5)
x→a x→a ′
g(x) g (x)

2. Let f and g be differentiable functions on the open interval (a, ∞) for some value a , where g ′
(x) ≠ 0 on (a, ∞) and
lim x→∞f (x)/g(x) returns either 0/0 or ∞/∞. Then


f (x) f (x)
lim = lim . (6.7.6)
x→∞ x→∞ ′
g(x) g (x)

A similar statement can be made for limits where x approaches −∞ .

Example 6.7.2: L'Hôpital's Rule with limits involving ∞

Evaluate the following limits.


$$ 1.\ \lim_{x\to\infty} \frac{3x^2-100x+2}{4x^2+5x-1000} \qquad\qquad 2. \ \lim_{x\to \infty}\frac{e^x}{x^3}.\]
Solution
1. We can evaluate this limit already using Theorem ??? ; the answer is 3/4. We apply L'Hôpital's Rule to demonstrate its
applicability.
2
3x − 100x + 2 by LHR 6x − 100 by LHR 6 3
lim = lim = lim = . (6.7.7)
2
x→∞ 4x + 5x − 1000 x→∞ 8x + 5 x→∞ 8 4

x x x x
2. e by LHR e by LHR e by LHR e
lim = lim = lim = lim = ∞. (6.7.8)
3 2
x→∞ x x→∞ 3x x→∞ 6x x→∞ 6

Recall that this means that the limit does not exist; as x approaches ∞, the expression e /x grows without bound. We can x 3

infer from this that e grows "faster" than x ; as x gets large, e is far larger than x . (This has important implications in
x 3 x 3

computing when considering efficiency of algorithms.)

L'Hôpital's Rule can only be applied to ratios of functions. When faced with an indeterminate form such as 0 ⋅ ∞ or ∞ − ∞ , we
can sometimes apply algebra to rewrite the limit so that L'Hôpital's Rule can be applied. We demonstrate the general idea in the
next example.

Example 6.7.3: Applying L'Hôpital's Rule to other indeterminate forms

Evaluate the following limits.


1. lim x→0
+ x⋅e
1/x

2. lim x→0
− x⋅e
1/x

3. lim x→∞ ln(x + 1) − ln x

4. lim x→∞ x
2
−e
x

Solution

6.7.2 https://math.libretexts.org/@go/page/4659
1/x

1. As x → 0 , x → 0 and e
+ 1/x
→ ∞ . Thus we have the indeterminate form 0 ⋅ ∞ . We rewrite the expression x ⋅ e 1/x
as e

1/x
;
now, as x → 0 , we get the indeterminate form ∞/∞ to which L'Hôpital's Rule can be applied.
+

1/x 2 1/x
e by LHR (−1/ x )e
1/x 1/x
lim x ⋅ e = lim = lim = lim e = ∞. (6.7.9)
+ + + 2 +
x→0 x→0 1/x x→0 −1/x x→0

Interpretation: e grows "faster" than x shrinks to zero, meaning their product grows without bound.
1/x

2. As x → 0 , x → 0 and e → e

→ 0 . The the limit evaluates to 0 ⋅ 0 which is not an indeterminate form. We
1/x −∞

conclude then that $$\lim_{x\to 0^-}x\cdot e^{1/x} = 0.\]


3. This limit initially evaluates to the indeterminate form ∞ − ∞ . By applying a logarithmic rule, we can rewrite the limit as
x +1
lim ln(x + 1) − ln x = lim ln( ). (6.7.10)
x→∞ x→∞ x

As x → ∞ , the argument of the ln term approaches ∞/∞, to which we can apply L'Hôpital's Rule.
x +1 by LHR 1
lim = = 1. (6.7.11)
x→∞ x 1

Since x → ∞ implies → 1 , it follows that $$x\rightarrow \infty \quad \text{ implies }\quad
x+1

\ln\left(\frac{x+1}x\right)\rightarrow \ln 1=0.\]


Thus
x +1
lim ln(x + 1) − ln x = lim ln( ) = 0. (6.7.12)
x→∞ x→∞ x

Interpretation: since this limit evaluates to 0, it means that for large x, there is essentially no difference between ln(x + 1)
and ln x; their difference is essentially 0.
4. The limit lim x→∞ x
2
−e
x
initially returns the indeterminate form ∞ − ∞ . We can rewrite the expression by factoring out
x

x
2
;x2
−e
x
=x
2
(1 −
e

x
2
). We need to evaluate how e x
/x
2
behaves as x → ∞ : $$\lim_{x\to\infty}\frac{e^x}{x^2}
\stackrel{\ \text{ by LHR } \ }{=} \lim_{x\to\infty} \frac{e^x}{2x} \stackrel{\ \text{ by LHR } \ }{=} \lim_{x\to\infty}
\frac{e^x}{2} = \infty.\]
Thus lim x (1 − e / x
x→∞
2 x 2
) evaluates to ∞ ⋅ (−∞) , which is not an indeterminate form; rather, ∞ ⋅ (−∞) evaluates to
−∞ . We conclude that lim
2 x
x→∞ x −e = −∞.

Interpretation: as x gets large, the difference between x and e grows very large. 2 x

Indeterminate Forms 0 , 1 0 ∞
and ∞ 0

When faced with an indeterminate form that involves a power, it often helps to employ the natural logarithmic function. The
following Key Idea expresses the concept, which is followed by an example that demonstrates its use.

Key Idea 20: Evaluating Limits Involving Indeterminate Forms 0 , 1 0 ∞


and ∞ 0

If lim x→c ln (f (x)) = L , then lim x→c f (x) = limx→c e


ln(f (x))
=e
L
.

Example 6.7.4: Using L'Hôpital's Rule with indeterminate forms involving exponents

Evaluate the following limits.


x
1 x
1. limx→∞ (1 + ) 2. limx→0+ x .
x

Solution
1. This equivalent to a special limit given in Theorem ??? ; 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) ; the problem asks to evaluate lim f (x). Let's first evaluate lim ln (f (x)).
x
x→∞ x→∞

6.7.3 https://math.libretexts.org/@go/page/4659
x
1
lim ln (f (x)) = lim ln (1 + ) (6.7.13)
x→∞ x→∞ x

1
= lim x ln(1 + ) (6.7.14)
x→∞ x
1
ln(1 + )
x
= lim (6.7.15)
x→∞ 1/x

This produces the indeterminate form 0/0, so we apply L'Hôpital's Rule.


1 2
⋅ (−1/ x )
1+1/x
= lim (6.7.16)
x→∞ 2
(−1/ x )

1
= lim (6.7.17)
x→∞ 1 + 1/x

= 1. (6.7.18)

Thus lim x→∞ ln (f (x)) = 1. We return to the original limit and apply Key Idea 20.

$$\lim_{x\to\infty}\left(1+\frac1x\right)^x = \lim_{x\to\infty} f(x) = \lim_{x\to\infty}e^{\ln (f(x))} = e^1 = e.\]


2. This limit leads to the indeterminate form 0 . Let f (x) = x and consider first lim
0 x
ln (f (x)). x→0
+

x
lim ln (f (x)) = lim ln(x ) (6.7.19)
+ +
x→0 x→0

= lim x ln x (6.7.20)
+
x→0

ln x
= lim . (6.7.21)
x→0
+
1/x

This produces the indeterminate form −∞/∞ so we apply L'Hôpital's Rule.


1/x
= lim (6.7.22)
+ 2
x→0 −1/x

= lim −x (6.7.23)
+
x→0

= 0. (6.7.24)

Thus lim x→0


+ ln (f (x)) = 0 . We return to the original limit and apply Key Idea 20.
x ln(f (x)) 0
lim x = lim f (x) = lim e =e = 1. (6.7.25)
+ + +
x→0 x→0 x→0

This result is supported by the graph of f (x) = x given in Figure 6.7.1.


x

Figure 6.7.1 : A graph of f (x) = x supporting the fact that as x → 0 , f (x) → 1.


x +

Our brief revisit of limits will be rewarded in the next section where we consider improper integration. So far, we have only
1
considered definite integrals where the bounds are finite numbers, such as ∫ f (x) dx . Improper integration considers integrals
0

6.7.4 https://math.libretexts.org/@go/page/4659
where one, or both, of the bounds are "infinity." Such integrals have many uses and applications, in addition to generating ideas that
are enlightening.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/
Integrated by Justin Marshall.

This page titled 6.7: L'Hopital's Rule is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et
al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

6.7.5 https://math.libretexts.org/@go/page/4659
6.8: Improper Integration
We begin this section by considering the following definite integrals:
100
1
∫ dx ≈ 1.5608, (6.8.1)
2
0 1 +x

1000
1
∫ dx ≈ 1.5698, (6.8.2)
2
0 1 +x

10,000
1
∫ dx ≈ 1.5707. (6.8.3)
2
0 1 +x

Notice how the integrand is 1/(1 + x ) in each integral (which is sketched in Figure 6.8.1). As the upper bound gets larger, one
2

would expect the "area under the curve" would also grow. While the definite integrals do increase in value as the upper bound
grows, they are not increasing by much. In fact, consider:
$$
b
1 b
−1
∫ dx = tan x∣
∣ (6.8.4)
2 0
0 1 +x

−1 −1
= tan b − tan 0 (6.8.5)

−1
= tan b. (6.8.6)

\]
b
As b → ∞ , −1
tan b → π/2. Therefore it seems that as the upper bound b grows, the value of the definite integral ∫
0
1

1+x2
dx

approaches π/2 ≈ 1.5708. This should strike the reader as being a bit amazing: even though the curve extends "to infinity," it has a
finite amount of area underneath it.

Figure 6.8.1 : Graphing f (x) = 1

1+x
2

b
When we defined the definite integral ∫ a
f (x) dx , we made two stipulations:
1. The interval over which we integrated, [a, b], was a finite interval, and
2. The function f (x) was continuous on [a, b] (ensuring that the range of f was finite).
In this section we consider integrals where one or both of the above conditions do not hold. Such integrals are called improper
integrals.

Improper Integrals with Infinite Bounds


Definition 6.8.1: Improper Integrals with Infinite Bounds; Converge, Diverge
1. Let f be a continuous function on [a, ∞). Define
∞ b

∫ f (x) dx ≡ lim ∫ f (x) dx. (6.8.7)


b→∞
a a

6.8.1 https://math.libretexts.org/@go/page/4660
2. Let f be a continuous function on (−∞, b] . Define
b b

∫ f (x) dx ≡ lim ∫ f (x) dx. (6.8.8)


a→−∞
−∞ a

3. Let f be a continuous function on (−∞, ∞). Let c be any real number; define
∞ c b

∫ f (x) dx ≡ lim ∫ f (x) dx + lim ∫ f (x) dx. (6.8.9)


a→−∞ b→∞
−∞ a c

An improper integral is said to converge if its corresponding limit exists; otherwise, it diverges. The improper integral in part
3 converges if and only if both of its limits exist.

Example 6.8.1: Evaluating improper integrals

Evaluate the following improper integrals.



1. ∫ 1
1

x
2
dx

2. ∫ 1
1

x
dx
0
3. ∫ −∞
e
x
dx

4. ∫ −∞ 1+x
1
2
dx

Solution

1. ∞
1
b
1 −1 b

[t] ∫ dx = lim ∫ dx = lim (6.8.10)

1
x2 b→∞
1
x2 b→∞ x 1

−1
= lim +1 (6.8.11)
b→∞ b

= 1. (6.8.12)

A graph of the area defined by this integral is given in Figure 6.8.2.

Figure 6.8.2 : A graph of f (x) = 1

x
2
in Example 6.8.1.

2. ∞
1
b
1
∫ dx = lim ∫ dx (6.8.13)
1
x b→∞
1
x

b

= lim ln |x| (6.8.14)

b→∞ 1

= lim ln(b) (6.8.15)


b→∞

= ∞. (6.8.16)


The limit does not exist, hence the improper integral ∫ dx diverges. Compare the graphs in Figures 6.8.3a and 6.8.3b;
1
1

notice how the graph of f (x) = 1/x is noticeably larger. This difference is enough to cause the improper integral to
diverge.

6.8.2 https://math.libretexts.org/@go/page/4660
Figure 6.8.3 : A graph of f (x) = 1

x
in Example 6.8.1

3. 0
x
0
x
∫ e dx = lim ∫ e dx (6.8.17)
a→−∞
−∞ a

0
x∣
= lim e (6.8.18)

a→−∞ a

0 a
= lim e −e (6.8.19)
a→−∞

= 1. (6.8.20)

A graph of the area defined by this integral is given in Figure 6.8.4.

Figure 6.8.4 : A graph of f (x) = e in Example 6.8.1 x

4. We will need to break this into two improper integrals and choose a value of c as in part 3 of Definition 6.8.1. Any value of
c is fine; we choose c = 0 .

∞ 0 b
1 1 1
∫ dx = lim ∫ dx + lim ∫ dx (6.8.21)
2 a→−∞ 2 2
−∞ 1 +x a 1 +x b→∞
0 1 +x
0 b
−1 ∣ −1 ∣
= lim tan x + lim tan x (6.8.22)
∣ ∣
a→−∞ a b→∞ 0

−1 −1 −1 −1
= lim (tan 0 − tan a) + lim (tan b − tan 0) (6.8.23)
a→−∞ b→∞

−π π
= (0 − ) +( − 0) . (6.8.24)
2 2

Each limit exists, hence the original integral converges and has value:

= π. (6.8.25)

A graph of the area defined by this integral is given in Figure 6.8.5.

6.8.3 https://math.libretexts.org/@go/page/4660
Figure 6.8.5 : A graph of f (x) = 1+x2
1
in Example 6.8.1

The previous section introduced L'Hôpital's Rule, a method of evaluating limits that return indeterminate forms. It is not
uncommon for the limits resulting from improper integrals to need this rule as demonstrated next.

Example 6.8.2: Improper integration and L'Hôpital's Rule

Evaluate the improper integral



ln x
∫ dx. (6.8.26)
2
1 x

Solution
This integral will require the use of Integration by Parts. Let u = ln x and dv = 1/x 2
dx . Then

Figure 6.8.6 : A graph of f (x) = ln x

x
2
in Example 6.8.2
∞ b
ln x ln x
∫ dx = lim ∫ dx (6.8.27)
2
1
x b→∞
1
x2
b
ln x b 1

= lim (− +∫ dx) (6.8.28)
∣1
b→∞ x 1 x2

b
ln x 1 ∣
= lim (− − )∣ (6.8.29)
b→∞ x x ∣1

ln b 1
= lim (− − − (− ln 1 − 1)) . (6.8.30)
b→∞ b b

The 1/b and ln 1 terms go to 0, leaving lim b→∞ −


ln b

b
+ 1. We need to evaluate lim b→∞
ln b

b
with L'Hôpital's Rule. We have:

ln b by LHR 1/b
lim = lim (6.8.31)
b→∞ b b→∞ 1

= 0. (6.8.32)

Thus the improper integral evaluates as:

6.8.4 https://math.libretexts.org/@go/page/4660

ln x
∫ dx = 1. (6.8.33)
2
1 x

Improper Integrals with Infinite Range


We have just considered definite integrals where the interval of integration was infinite. We now consider another type of improper
integration, where the range of the integrand is infinite.

Definition 6.8.2: Improper Integration with Infinite Range

{Let f (x) be a continuous function on [a, b] except at c , a ≤ c ≤ b , where x = c is a vertical asymptote of f . Define
b t b

∫ f (x) dx = lim ∫ f (x) dx + lim ∫ f (x) dx. (6.8.34)


− +
t→c t→c
a a t

Example 6.8.3: Improper integration of functions with infinite range

Evaluate the following improper integrals:


1 1 1 1
1. ∫ dx 2. ∫ dx.
0 √x −1 2
x

Solution
1. A graph of f (x) = 1/√− x is given in Figure 6.8.7. Notice that f has a vertical asymptote at x = 0 ; in some sense, we are

trying to compute the area of a region that has no "top." Could this have a finite value?
1 1
1 1
∫ dx = lim ∫ dx (6.8.35)
− +

0 √x a→0 a √x
1
−∣
= lim 2 √x (6.8.36)
+

a→0 a

– −

= lim 2 (√1 − √a ) (6.8.37)
+
a→0

= 2. (6.8.38)

It turns out that the region does have a finite area even though it has no upper bound (strange things can occur in
mathematics when considering the infinite).
Note: In Definition 6.8.1, c can be one of the endpoints (a or b ). In that case, there is only one limit to consider as part of
the definition.

Figure 6.8.7 : A graph of f (x) = √x


1
in Example 6.8.3

2. The function f (x) = 1/x has a vertical asymptote at x = 0 , as shown in Figure 6.8.8, so this integral is an improper
2

integral. Let's eschew using limits for a moment and proceed without recognizing the improper nature of the integral. This
leads to:

6.8.5 https://math.libretexts.org/@go/page/4660
1
1 1 1

∫ dx = − (6.8.39)
2 ∣−1
−1 x x

= −1 − (1) (6.8.40)

= −2! (6.8.41)

Clearly the area in question is above the x-axis, yet the area is supposedly negative! Why does our answer not match our
intuition? To answer this, evaluate the integral using Definition 6.8.2.
1 t 1
1 1 1
∫ dx = lim ∫ dx + lim ∫ dx (6.8.42)
2 − 2 + 2
−1 x t→0 −1 x t→0 t x

1 t 1 1
∣ ∣
= lim − + lim − (6.8.43)
∣ ∣
t→0

x −1 t→0
+
x t

1 1
= lim − − 1 + lim −1 + (6.8.44)
t→0

t t→0
+
t

⇒ (∞ − 1) + ( − 1 + ∞). (6.8.45)

Neither limit converges hence the original improper integral diverges. The nonsensical answer we obtained by ignoring the
improper nature of the integral is just that: nonsensical.

Figure 6.8.8 : A graph of f (x) = x2


1
in Example 6.8.3

Understanding Convergence and Divergence


Oftentimes we are interested in knowing simply whether or not an improper integral converges, and not necessarily the value of a
convergent integral. We provide here several tools that help determine the convergence or divergence of improper integrals without
integrating.
Our first tool is to understand the behavior of functions of the form 1

x
p
.

Example 6.8.4: Improper integration of 1/x p


Determine the values of p for which ∫ 1
1

xp
dx converges.
Solution
We begin by integrating and then evaluating the limit.
∞ b
1 1
∫ dx = lim ∫ dx (6.8.46)
p p
1
x b→∞
1
x
b
−p
= lim ∫ x dx (assume p ≠ 1) (6.8.47)
b→∞
1

1 b
−p+1 ∣
= lim x (6.8.48)
∣1
b→∞ −p + 1

1 1−p 1−p
= lim (b −1 ). (6.8.49)
b→∞ 1 −p

6.8.6 https://math.libretexts.org/@go/page/4660
When does this limit converge -- i.e., when is this limit not ∞? This limit converges precisely when the power of b is less than
0: when 1 − p < 0 ⇒ 1 < p .

Figure 6.8.9 : Plotting functions of the form 1/x in Example 6.8.4 p


Our analysis shows that if p > 1 , then ∫ dx converges. When
1
1

x
p
p <1 the improper integral diverges; we showed in
Example 6.8.1 that when p = 1 the integral also diverges.
Figure 6.8.9 graphs y = 1/x with a dashed line, along with graphs of y = 1/x , p < 1 , and p
y = 1/x
q
, q >1 . Somehow the
dashed line forms a dividing line between convergence and divergence.

The result of Example 6.8.4 provides an important tool in determining the convergence of other integrals. A similar result is proved
1
in the exercises about improper integrals of the form ∫ 0
1

x
p dx . These results are summarized in the following Key Idea.
∞ 1
Key Idea 21: Convergence of Improper Integrals ∫ 1
1

x
p dx and ∫ 0
1

x
p dx .

1. The improper integral ∫ 1 x
1
p
dx converges when p > 1 and diverges when p ≤ 1.
1
2. The improper integral ∫ 0
1

x
p
dx converges when p < 1 and diverges when p ≥ 1.

A basic technique in determining convergence of improper integrals is to compare an integrand whose convergence is unknown to
an integrand whose convergence is known. We often use integrands of the form 1/x to compare to as their convergence on certain p

intervals is known. This is described in the following theorem.

Theorem 6.8.1: Direct Comparison Test for Improper Integrals

Let f and g be continuous on [a, ∞) where 0 ≤ f (x) ≤ g(x) for all x in [a, ∞).
∞ ∞
1. If ∫ a
g(x) dx converges, then ∫ a
f (x) dx converges.
∞ ∞
2. If ∫ a
f (x) dx diverges, then ∫
a
g(x) dx diverges.

Note: We used the upper and lower bound of "1" in Key Idea 21 for convenience. It can be replaced by any a where a > 0 .

Example 6.8.5: Determining convergence of improper integrals

Determine the convergence of the following improper integrals.



1. ∫
2
−x
e dx
1

2. ∫3
1
dx
√x2 −x

Solution
2

1. The function f (x) = e does not have an antiderivative expressible in terms of elementary functions, so we cannot
−x

integrate directly. It is comparable to g(x) = 1/x , and as demonstrated in Figure 6.8.10, e < 1/ x on [1, ∞). We
2
2 −x 2

∞ ∞ 2

know from Key Idea 21 that ∫ dx converges, hence ∫


1 x
1
2
e dx also converges.
1
−x

6.8.7 https://math.libretexts.org/@go/page/4660
Figure 6.8.10 : Graphs of f (x) = e and f (x) = 1/x in Example 6.8.6
2
−x 2


2. Note that for large values of x, 1

1
2
=
1

x
. We know from Key Idea 21 and the subsequent note that ∫ 3
1

x
dx
√x2 −x √x

diverges, so we seek to compare the original integrand to 1/x.


−− −−−− −
It is easy to see that when x > 0 , we have x = √x > √x − x . Taking reciprocals reverses the inequality, giving
2 2

1 1
< −−−−−. (6.8.50)
x 2
√x − x

∞ ∞
Using Theorem 6.8.1, we conclude that since ∫ 3
1

x
dx diverges, ∫ 3
1
dx diverges as well. Figure 6.8.11 illustrates
√x2 −x

this.

−−−−−
Figure 6.8.11 : Graphs of f (x) = 1/√x 2
−x and f (x) = 1/x in Example 6.8.5

Being able to compare "unknown" integrals to "known" integrals is very useful in determining convergence. However, some of our
examples were a little "too nice." For instance, it was convenient that < , but what if the "−x" were replaced with a "
1

x
1

√x2 −x

+2x + 5 "? That is, what can we say about the convergence of ∫
3
1
dx ? We have 1

x
>
1
, so we cannot use
√x2 +2x+5 √x2 +2x+5

Theorem 6.8.1.
In cases like this (and many more) it is useful to employ the following theorem.

Theorem: Limit Comparison Test for Improper Integrals


Let f and g be continuous functions on [a, ∞) where f (x) > 0 and g(x) > 0 for all x. If
f (x)
lim = L, 0 < L < ∞, (6.8.51)
x→∞ g(x)

then
∞ ∞

∫ f (x) dx and ∫ g(x) dx (6.8.52)


a a

either both converge or both diverge.

6.8.8 https://math.libretexts.org/@go/page/4660
Example 6.8.6: Determining convergence of improper integrals

Determine the convergence of ∫ 3
1
dx .
√x2 +2x+5

Solution
As x gets large, the quadratic inside the square root function will begin to behave much like y = x . So we compare 1
\
√x2 +2x+5

to 1

x
with the Limit Comparison Test:
$$\lim_{x\to\infty} \frac{1/\sqrt{x^2+2x+5}}{1/x} = \lim_{x\to\infty}\frac{x}{\sqrt{x^2+2x+5}}.\]
The immediate evaluation of this limit returns ∞/∞, an indeterminate form. Using L'Hôpital's Rule seems appropriate, but in
this situation, it does not lead to useful results. (We encourage the reader to employ L'Hôpital's Rule at least once to verify
this.)
The trouble is the square root function. To get rid of it, we employ the following fact: If limx→c f (x) = L , then
f (x ) = L . (This is true when either c or L is ∞.) So we consider now the limit\)
2 2
lim x→c

$$\lim_{x\to\infty} \frac{x^2}{x^2+2x+5}.\]
This converges to 1, meaning the original limit also converged to 1. As x gets very large, the function 1
looks very
√x2 +2x+5

∞ ∞
much like 1

x
. Since we know that ∫
3
1

x
dx diverges, by the Limit Comparison Test we know that ∫
3
1
dx also
√x2 +2x+5
−−−−−−−− −
diverges. Figure 6.8.12 graphs 2
f (x) = 1/ √x + 2x + 5 and f (x) = 1/x , illustrating that as x gets large, the functions
become indistinguishable.

Figure 6.8.12 : Graphing f (x) = 1

√x2 +2x+5
and f (x) = 1

x
in Example 6.8.6.

Both the Direct and Limit Comparison Tests were given in terms of integrals over an infinite interval. There are versions that apply
to improper integrals with an infinite range, but as they are a bit wordy and a little more difficult to employ, they are omitted from
this text.
This chapter has explored many integration techniques. We learned Substitution, which "undoes" the Chain Rule of differentiation,
as well as Integration by Parts, which "undoes" the Product Rule. We learned specialized techniques for handling trigonometric
functions and introduced the hyperbolic functions, which are closely related to the trigonometric functions. All techniques
effectively have this goal in common: rewrite the integrand in a new way so that the integration step is easier to see and implement.
As stated before, integration is, in general, hard. It is easy to write a function whose antiderivative is impossible to write in terms of
elementary functions, and even when a function does have an antiderivative expressible by elementary functions, it may be really
hard to discover what it is. The powerful computer algebra system Mathematica has approximately 1,000 pages of code dedicated
to integration.
Do not let this difficulty discourage you. There is great value in learning integration techniques, as they allow one to manipulate an
integral in ways that can illuminate a concept for greater understanding. There is also great value in understanding the need for
good numerical techniques: the Trapezoidal and Simpson's Rules are just the beginning of powerful techniques for approximating
the value of integration.
The next chapter stresses the uses of integration. We generally do not find antiderivatives for antiderivative's sake, but rather
because they provide the solution to some type of problem. The following chapter introduces us to a number of different problems

6.8.9 https://math.libretexts.org/@go/page/4660
whose solution is provided by integration.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/
Integrated by Justin Marshall.

This page titled 6.8: Improper Integration is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory
Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

6.8.10 https://math.libretexts.org/@go/page/4660
6.E: Applications of Antidifferentiation (Exercises)
6.1: Substitution
Terms and Concepts
1. Substitution "undoes" what derivative Rule?
2. T/F: One can use algebra to rewrite the integrand of an integral to make it easier to evaluate.

Problems
In Exercises 3-14, evaluate the indefinite integrand to develop an understanding of Substitution.
3. ∫ 3x 2
(x
3
− 5)
7
dx

4. ∫ (2x − 5)(x 2
− 5x + 7 )
3
dx

5. ∫ x(x 2
+ 1)
8
dx

6. ∫ (12x + 14)(3x 2
+ 7x + 7 )
3
dx

7. ∫ 1

2x+7
dx

8. ∫ 1
dx
√2x+3

9. ∫ x

√x+3
dx

10. ∫ x −x

√x
dx

√x

11. ∫ e
dx
√x

12. ∫ x
dx
√x5 +1

1
+1
13. ∫ x

x
2
dx

ln(x)
14. ∫ x
dx

In Exercises 15-23, use Substitution to evaluate the indefinite integral involving trigonometric functions.
15. ∫ sin 2
(x) cos(x) dx

16. ∫ cos(3 − 6x) dx


17. ∫ sec 2
(4 − x) dx

18. ∫ sec(2x) dx
19. ∫ tan 2
(x) sec (x) dx
2

20. ∫ x cos(x 2
) dx

21. ∫ tan 2
(x) dx

22. ∫ cot x dx. Do not just refer to Theorem 45 for the answer; justify it through Substitution.
23. ∫ csc x dx. Do not just refer to Theorem 45 for the answer; justify it through Substitution.
In Exercises 24-30, use Substitution to evaluate the indefinite integral involving exponential functions.
24. ∫ e 3x−1
dx

25. ∫ e
3
x 2
x dx

26. ∫ e x −2x+1
(x − 1) dx

27. ∫ e +1

e
x
dx

x −x
e −e
28. ∫ e
2x
dx

6.E.1 https://math.libretexts.org/@go/page/9976
29. ∫ 3 3x
dx

30. ∫ 4 2x
dx

In Exercises 31-34, use Substitution to evaluate the indefinite integral involving logarithmic functions.
31. ∫ ln x

x
dx

2
(ln x)
32. ∫ x
dx

3
(ln x)
33. ∫ x
dx

34. ∫ x ln( x2 )
1
dx

In Exercises 35-40, use Substitution to evaluate the indefinite integral involving rational functions.
2

35. ∫ x +3x+1

x
dx

3 2

36. ∫ x +x +x+1

x
dx

37. ∫ x −1

x+!
dx

2
x +2x−5
38. ∫ x−3
dx

2
3 x −5x+7
39. ∫ x+1
dx

40. ∫ x +2x+1

x3 +3 x2 +3x
dx

In Exercises 41-50, use Substitution to evaluate the indefinite integral inverse trigonometric functions.
41. ∫ 7

x2 +7
dx

42. ∫ 3
dx
√9−x2

43. ∫ 14
dx
√5−x2

44. ∫ 2
dx
x√x2 −9

45. ∫ 5
dx
√x4 −16 x2

46. ∫ x
dx
√1−x4

47. ∫ 2
x −2x+8
1
dx

48. ∫ 2
dx
√−x2 +6x+7

49. ∫ 3
dx
√−x2 +8x+9

50. ∫ x2 +6x+34
5
dx

In Exercises 51-75, evaluate the indefinite integral.


2

51. ∫ x
2
dx
( x3 +3 )

52. ∫ (3x 2
+ 2x)(5 x
3 2
+ 5x
8
+ 2) dx

53. ∫ x
dx
√1−x2

54. ∫ x 2
csc (x
2 3
+ 1) dx

−−−−−
55. ∫ sin(x)√cos(x) dx

56. ∫ 1

x−5
dx

57. ∫ 3x+2
7
dx

6.E.2 https://math.libretexts.org/@go/page/9976
3 2

58. ∫ 3 x +4 x +2x−22
2
x +3x+5
dx

59. ∫ 2x+7
2
x +7x+3
dx

9(2x+3)
60. ∫ 2
3 x +9x+7
dx

3 2

61. ∫ −x +14 x −46x−7

x2 −7x+1
dx

62. ∫ 2
x +81
x
dx

63. ∫ 4 x2 +1
2
dx

64. ∫ 1
dx
x√4 x2 −1

65. ∫ 1
dx
√16−9x2

66. ∫ 2
3x−2

x −2x+10
dx

67. ∫ 7−2x

x2 +12x+61
dx

68. ∫ x +5x−2
2
x −10x+32
dx

69. ∫ x
2
x +9
dx

70. ∫ x −x
2
x +4x+9
dx

sin(x)
71. ∫ cos2 (x)+1
dx

cos(x)
72. ∫ 2
dx
sin (x)+1

cos(x)
73. ∫ 2
dx
1−sin (x)

74. ∫ 3x−3
dx
√x2 −2x−6

75. ∫ x−3
dx
√x2 −6x+8

In Exercises 76-83, evaluate the definite integral.


3
76. ∫1 x−5
1
dx

6 −−−−−
77. ∫2
x √x − 2 dx

π/2
78. ∫−π/2
2
sin (x) cos(x) dx

1
79. ∫0
2x(1 − x )
2 4
dx

−1
80. ∫
2
x +2x+1
(x + 1)e dx
−2

1
81. ∫−1 x+x
1
2
dx

4
82. ∫2 2
x −6x+10
1
dx

√3
83. ∫1
1
dx
√4−x2

6.2: Integration by Parts


Terms and Concepts
1. T/F: Integration by Parts is useful in evaluating integrands that contain products of function.
2. T/F: Integration by Parts can be thought of as the "opposite of the Chain Rule."
3. For what is "LIATE" useful?

6.E.3 https://math.libretexts.org/@go/page/9976
Problems
In Exercises 4-33, evaluate the given indefinite integral.
4. ∫ x sin x dx
5. ∫ x e −x
dx

6. ∫ x 2
sin x dx

7. ∫ x 3
sin x dx

8. ∫ x e
2
x
dx

9. ∫ x 3
e
x
dx

10. ∫ x e −2x
dx

11. ∫ e x
sin x dx

12. ∫ e 2x
cos x dx

13. ∫ e 2x
sin(3x) dx

14. ∫ e 5x
cos(5x) dx

15. ∫ sin x cos x dx


16. ∫ sin −1
x dx

17. ∫ tan −1
(2x) dx

18. ∫ x tan −1
x dx

19. ∫ sin −1
x dx

20. ∫ x ln x dx
21. ∫ (x − 2) ln x dx
22. ∫ x ln(x − 1) dx
23. ∫ x ln(x 2
) dx

24. ∫ x 2
ln x dx

25. ∫ (ln x ) 2
dx

26. ∫ (ln(x + 1)) 2


dx

27. ∫ x sec 2
x dx

28. ∫ x csc 2
x dx

−−−−−
29. ∫ x √x − 2 dx

−−−−−
30. ∫ x √x 2
− 2 dx

31. ∫ sec x tan x dx


32. ∫ x sec x tan x dx
33. ∫ x csc x cot x dx
In Exercises 34-38, evaluate the indefinite integral after first making a substitution.
34. ∫ sin(ln x) dx
35. ∫ sin(√−
x ) dx

36. ∫ ln(√−
x ) dx

37. ∫ e √x
dx

38. ∫ e ln x
dx

6.E.4 https://math.libretexts.org/@go/page/9976
In Exercises 39-47, evaluate the definite integral. Note: the corresponding indefinite integrals appear in Exercises 4-12.
π
39. ∫0
x sin x dx

1
40. ∫−1
xe
−x
dx

41. ∫−π/4
π
/4 x
2
sin x dx

π/2
42. ∫−π/2
x
3
sin x dx

√ln 2 2

43. ∫0
xe
x
dx

1
44. ∫0
x e
3 x
dx

2
45. ∫1
xe
−2x
dx

π
46. ∫0
e
x
sin x dx

π/2
47. ∫−π/2
e
2x
cos x dx

6.3: Trigonometric Integrals


Terms and Concepts
1. T/F: ∫ sin 2
(x) cos
2
x dx cannot be evaluate using the techniques described in this section since both powers of sin x and cos x
are even.
2. T/F: sin 3
x cos
3
x dx cannot be evaluated using the techniques described in this section since both powers of sin x and cos x

are odd.
3. T/F: This section addresses how to evaluate indefinite integrals such as ∫ sin 5 3
x tan x dx .

Problems
In Exercises 4-26, evaluate the indefinite integral.
4. ∫ sin x cos 4
x dx

5. ∫ sin 3
x cos x dx

6. ∫ sin 3
x cos
2
x dx

7. ∫ sin 3
x cos
3
x dx

8. ∫ sin 6
x cos
5
x dx

9. ∫ sin 2
x cos
7
x dx

10. ∫ sin 2
x cos
2
x dx

11. ∫ sin(5x) cos(3x) dx


12. ∫ sin(x) cos(2x) dx
13. ∫ sin(3x) sin(7x) dx
14. ∫ sin(πx) sin(2πx) dx
15. ∫ cos(x) cos(2x) dx
16. ∫ cos( π

2
x) cos(πx) dx

17. ∫ tan 4
x sec
2
x dx

18. ∫ tan 2
x sec
4
x dx

19. ∫ tan 3
x sec
4
x dx

20. ∫ tan 3
x sec
2
x dx

21. ∫ tan 3
x sec
3
x dx

6.E.5 https://math.libretexts.org/@go/page/9976
22. ∫ tan 5
x sec
5
x dx

23. ∫ tan 4
(x) dx

24. ∫ sec 5
x dx

25. ∫ tan 2
x sec x dx

26. ∫ tan 2
x sec
3
x dx

In Exercises 27-33, evaluate the definite integral. Note: the corresponding indefinite integrals appear in the previous set.
π
27. ∫ 0
sin x cos
4
x dx

π
28. ∫ −π
sin
3
x cos x dx

π/2
29. ∫ −π/2
sin
2
x cos
7
x dx

π/2
30. ∫ 0
sin(5x) cos(3x) dx

π/2
31. ∫ −π/2
cos(x) cos(2x) dx

π/4
32. ∫ 0
tan
4
x sec
2
x dx

π/4
33. ∫ −π/4
tan
2
x sec
4
x dx

6.4: Trigonometric Substitution


Terms and Concepts
1. Trigonometric Substitution works on the same principles as Integration by Substitution, though it can feel "_____".
−−−−−−
2. If one uses Trigonometric Substitution on an integrand containing √25 − x , then one should set x = ______.
2

3. Consider the Pythagorean Identity sin θ + cos θ = 1 .


2 2

(a) What identity is obtained when both sides are divided by cos 2
θ ?
(b) Use the new identity to simplify 9 tan θ + 9 . 2

−− −−−−
4. Why does Key Idea 13(a) state that √a 2
−x
2
= a cos θ , and not |a cos θ| ?

Problems
In Exercises 5-16, apply Trigonometric Substitution to evaluate the indefinite integrals.
−−−−−
5. ∫ √x 2
+ 1 dx

−−−−−
6. ∫ √x 2
+ 4 dx

−−−−−
7. ∫ √1 − x 2
dx

−−−− −
8. ∫ √9 − x2 dx

−−−−−
9. ∫ √x 2
− 1 dx

−−−−− −
10. ∫ √x 2
− 16 dx

−−−−−−
11. ∫ √4x 2
+ 1 dx

−−−−−−
12. ∫ √1 − 9x 2
dx

−−− − −−−
13. 2
∫ √16 x − 1 dx

14. ∫ 3
dx
√x2 +2

15. ∫ 3
dx
√7−x2

16. ∫ 5
dx
√x2 −8

In Exercises 17-26, evaluate the indefinite integrals. Some may be evaluated without Trigonometric Substitution.

6.E.6 https://math.libretexts.org/@go/page/9976
√x2 −11
17. ∫ x
dx

18. ∫ 2
1
2
dx
( x +1 )

19. ∫ x
dx
√x2 −3

−−−− −
20. ∫ x 2
√1 − x2 dx

21. ∫ x

3/2
dx
( x2 +0 )

22. ∫ 5x
dx
√x2 −10

23. ∫ 2
1
2
dx
( x +4x+13 )

24. ∫ x 2
(1 − x )
2 −3/2
dx

√5−x2
25. ∫ 7x
2
dx

26. ∫ x
dx
√x2 +3

In Exercises 27-32, evaluate the definite integrals by making the proper trigonometric substitution and changing the
bounds of integration. (Note: each of the corresponding indefinite integrals has appeared previously in the Exercise set.)
1 −−−− −
27. ∫ −1
√1 − x2 dx

8 −−−−− −
28. ∫ 4
√x2 − 16 dx

2 −−−−−
29. ∫ 0
√x2 + 4 dx

1
30. ∫ −1 2
1
2
dx
( x +1 )

1 −−−
31. ∫ −1
√9x2 dx

1 −−−− −
32. ∫ −1
2 2
x √1 − x dx

6.5 Partial Fraction Decomposition


Terms and Concepts
1. Fill in the blank: Partial Fraction Decomposition is a method of rewriting _____ functions.
2. T/F: It is sometimes necessary to use polynomial division before using Partial Fraction Decomposition.
3. Decompose 2
x −3x
1
without solving for the coefficients, as done in Example 181.

4. Decompose x −9
7−x
2
without solving for the coefficients, as done in Example 181.

5. Decompose x −7
x−3
2
without solving for the coefficients, as done in Example 181.

6. Decompose 2x+5
3
x +7x
without solving for the coefficients, as done in Example 181.

Problems
In Exercises 7-25, evaluate the indefinite integral.
7. ∫ 7x+7

x2 +3x−10
dx

7x−2
8. ∫ x +x
2
dx

9. ∫ 3 x −12
−4
2
dx

10. ∫ x+7
2
dx
(x+5)

−3x−20
11. ∫ 2
dx
(x+8)

6.E.7 https://math.libretexts.org/@go/page/9976
2

12. ∫ 9 x +11x+7
2
dx
x(x+1)

13. ∫ −12 x −x+33

(x−1)(x+3)(3−2x)
dx

14. ∫ 94 x −10x

(7x+3)(5x−1)(3x−1)
dx

15. ∫ x +2+1

x2 +x−2
dx

16. ∫ x2 −2x−20
x
dx

17. ∫ 2 x −4x+6
2
x −2x+3
dx

18. ∫ x2 +3 x2 +3x
1
dx

19. ∫ x +x+5
2
x +4x+10
dx

20. ∫ 12 x +21x+3

(x+1)(3 x +5x−1)
2
dx

21. ∫ 6 x +8x−4

(x−3)( x +6x+10)
2
dx

22. ∫ 2 x +x+1

(x+1)( x2 +9)
dx

2
x −20x−69
23. ∫ (x−7)( x +2x+17)
2
dx

24. ∫ 9 x −60x+33

(x−9)( x −2x+11)
2
dx

25. ∫ 6 x +45x+121

(x+2)( x2 +10x+27)
dx

In Exercises 26-29, evaluate the definite integral.


2
26. ∫ 1
8x+21

(x+2)(x+3)
dx

5 14x+6
27. ∫ 0 (3x+2)(x+4)
dx

2
1
28. ∫ −1 (x−10)( x +4x+5)
x +5x−5
2
dx

1
29. ∫ 0 (x+1)( x2 +2x+1)
x
dx

6.6: Hyperbolic Functions


Terms and Concepts
1. In Key Idea 16, the equation ∫ tanh x dx = ln(cosh x) + C is given. Why is "ln | cosh x|" not used -i.e., why are absolute
values no necessary?
2. The hyperbolic functions are used to define points on the right hand portion of the hyperbola x − y 2 2
=1 , as shown in Figure
6.13. How can we use the hyperbolic functions to define points on the left hand portion of the hyperbola?

Problems
In Exercises 3-10, verify the given identity using Definition 23, as done in Example 186.
3. coth 2
x − csch
2
x =1

4. cosh 2x = cosh 2
x + sinh
2
x

5. cosh 2
x =
cosh 2x+1

6. sinh 2
x =
cosh 2x−1

7. d

dx
[sech x] = −sech x tanh x

8. d

dx
[coth x] = −sech x tanh x

6.E.8 https://math.libretexts.org/@go/page/9976
9. ∫ tanh x dx = ln(cosh x) + C
10. ∫ coth x dx = ln | sinh x| + C
In Exercises 11-21, find the derivative of the given function.
11. f (x) = cosh 2x
12. f (x) = tanh(x 2
)

13. f (x) = ln(sinh x)


14. f (x) = sinh x cosh x
15. f (x) = x sinh x − cosh x
16. f (x) = sech −1
(x )
2

17. f (x) = sinh −1


(3x)

18. f (x) = cosh −1


(2 x )
2

19. f (x) = tanh −1


(x + 5)

20. f (x) = tanh −1


(cos x)

21. f (x) = cosh −1


(sec x)

In Exercises 22-26, find the equation of the line tangent to the function at the given x-value.
22. f (x) = sinh x at x = 0
23. f (x) = cosh x at x = ln 2
24. f (x) = sech 2
x at x = ln 3

25. f (x) = sinh −1


x at x = 0


26. f (x) = cosh −1
x at x = √2

In Exercises 27-40, evaluate the given indefinite integral.


27. ∫ tanh(2x) dx
28. ∫ cosh(3x − 7) dx
29. ∫ sinh x cosh x dx
30. ∫ x cosh x dx
31. ∫ x sinh x dx
32. ∫ 9−x
1
2
dx

33. ∫ 2x
dx
√x4 −4

√x
34. ∫ dx
√1+x3

35. ∫ x2 −16
1
dx

36. ∫ x +x
2
1
dx

37. ∫ x
2x
e

+1
dx

38. ∫ sinh −1
x dx

39. ∫ tanh −1
x dx

40. ∫ sech x dx (Hint: multiply by cosh x

cosh x
; set u = sinh x .)
In Exercises 41-43, evaluate the given definite integral.

6.E.9 https://math.libretexts.org/@go/page/9976
1
41. ∫ −1
sinh x dx

ln 2
42. ∫ − ln 2
cosh x dx

1
43. ∫ 0
tanh
−1
.
x dx

6.7: L'Hopital's Rule


Terms and Concepts
1. List the different indeterminate forms described in this section.
2. T/F: l'Hopital's Rule provides a faster method of computing derivatives.

f (x) f (x)
3. T/F: l'Hopitals Rule states that d

dx
(
g(x)
) =
g ′ (x)
.

4. Explain what the indeterminate form "1 " means. ∞

f (x)
5. Fill in the blanks" The Quotient Rule is applied to g(x)
when taking _____; l'Hopital's Rule is applied when taking
certain_______.
6. Create (but do not evaluate) a limit that returns "∞ ". 0

7. Create a function f (x) such that lim f (x) returns "0 ". 0

x→1

Problems
In Exercises 8-52, evaluate the given limit.
2

8. lim x +x−2

x−1
x→1

9. lim x +x−6
2
x −7x+10
x→2

10. lim
sin x

x−π
x→π

11. lim
sin x−cos x

cos(2x)
x→π/4

sin(5x)
12. lim x
x→0

sin(2x)
13. lim x+2
x→0

sin(2x)
14. lim sin(3x)
x→0

sin(ax)
15. lim sin(bx)
x→0

16. lim
+
e −1

x
2

x→0

x
e −x−1
17. lim
+ x
2

x→0

18. lim
x−sin x

x3 −x2
+
x→0

19. lim
x

e
x
x→∞

√x
20. lim
e
x
x→∞

21. lim
e

√x
x→∞

22. lim
e

2
x
x→∞

23. lim
e

3
x
x→∞

6.E.10 https://math.libretexts.org/@go/page/9976
3 2

24. lim x −5 x +3x+9


3
x −7 x +15x−9
2
x→3

3 2

25. lim 3
x +4 x +4x

x +7 x +16x+12
2
x→−2

26. lim
ln x

x
x→∞

2
ln( x )
27. lim
x
x→∞

2
(ln x)
28. lim
x
x→∞

29. lim x ⋅ ln x
+
x→0


30. lim √x ⋅ ln x
+
x→0

31. lim x e
+
1/x

x→0

32. lim x
3
−x
2

x→∞


33. lim √x − ln x
x→∞

34. lim xe
x

x→−∞

35. lim
+
1

x
2
e
−1/x

x→0

36. lim (1 + x )
+
1/x

x→0

37. lim (2x )


x

x→0+

38. lim (2/x )


+
x

x→0

39. lim (sin x )


+
x
Hint: use the Squeeze Theorem.
x→0

40. lim (1 − x )
+
1−x

x→1

41. lim (x )
1/x

x→∞

42. lim (1/x )


x

x→∞

43. lim (ln x )


1
1−x

x→1

44. lim (1 + x )
1/x

x→∞

45. lim (1 + x )
2 1/x

x→∞

46. lim tan x cos x


x→π/2

47. lim tan x sin(2x)


x→π/2

48. lim
+ ln x
1

1−x
1

x→1

49. lim
+ x −9
2
5

x−3
x

x→3

50. lim x tan(1/x)


x→∞

3
(ln x)
51. lim
x
x→∞

2
x +x−2
52. lim ln x
x→1

6.E.11 https://math.libretexts.org/@go/page/9976
6.8: Improper Integration
Terms and Concepts
1. The definite integral was defined with what two stipulations?
b ∞
2. If lim ∫
0
f (x) dx exists, then the integral ∫ 0
f (x) dx is said to __________.
b→∞

∞ ∞
3. If ∫ 1
f (x) dx = 10, and 0 ≤ g(x) ≤ f (x) for all x, then we know that ∫1
g(x) dx ______.

4. For what values of p will ∫ 1 x
1
p dx converge?

5. For what values of p will ∫ 10 xp
1
dx converge?
1
6. For what values of p will ∫ 0
1

xp
dx converge?

Problems
In Exercises 7-33, evaluate the given improper integral.

7. ∫0
e
5−2x
dx


8. ∫1
1

x
3
dx


9. ∫1
x
−4
dx


10. ∫ −∞ x2 +9
1
dx

0
11. ∫ −∞
2
x
dx

0 x
12. ∫ −∞
(
1

2
) dx


13. ∫ −∞ x2 +1
x
dx


14. ∫ −∞ x +4
2
x
dx


15. ∫ 2
1
2
dx
(x−1)

2
16. ∫ 1
1
2
dx
(x−1)


17. ∫ 2 x−1
1
dx

2
18. ∫ 1 x−1
1
dx

1
19. ∫ −1
1

x
dx

3
20. ∫ 1 x−2
1
dx

π
21. ∫ 0
sec
2
x dx

1
22. ∫ −2
1
dx
√|x|


23. ∫ 0
xe
−x
dx

2

24. ∫ 0
xe
−x
dx


25. ∫
2
−x
xe dx
−∞


26. ∫ −∞ ex +e−x
1
dx

1
27. ∫ 0
x ln x dx


28. ∫ 1
ln x

x
dx

1
29. ∫ 0
ln x dx


30. ∫ 1
ln x

x
2
dx

6.E.12 https://math.libretexts.org/@go/page/9976

31. ∫1
ln x

√x
dx


32. ∫0
e
−x
sin x dx


33. ∫0
e
−x
cos x dx

In Exercises 34-43, use the Direct Comparison Test or the Limit Comparison Test to determine whether the given definite
integral converges or diverges. Clearly state what test is being used and what function the integrand is being compared to.

34. ∫10
3
dx
√3 x2 +2x−5


35. ∫2
4
dx
√7 x3 −x

∞ √x+3
36. ∫0
dx
√x3 −x2 +x+1


37. ∫1
e
−x
ln x dx


38. ∫
2
−x +3x−1
e dx
5

∞ √x
39. ∫0 e
x
dx


40. ∫2 x2 +sin x
1
dx


41. ∫0 2
x +cos x
x
dx


42. ∫0 x+e
1
x dx


43. ∫0 e −x
x
1
dx

6.E: Applications of Antidifferentiation (Exercises) is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by
LibreTexts.

6.E.13 https://math.libretexts.org/@go/page/9976
CHAPTER OVERVIEW
Back Matter
Index

1 https://math.libretexts.org/@go/page/45270
Index
A curvature F
absolute convergence 11.5: The Arc Length Parameter and Curvature factorial
8.5: Alternating Series and Absolute Convergence curve sketching 8.1: Sequences
acceleration 3.5: Curve Sketching FINDING POSITION USING
11.3: The Calculus of Motion cusps
VELOCITY
Alternating Harmonic Series 9.2: Parametric Equations
5.2: The Definite Integral
8.5: Alternating Series and Absolute Convergence cylinders
First Derivative Test
alternating series 10.1: Introduction to Cartesian Coordinates in Space
3.3: Increasing and Decreasing Functions
8.5: Alternating Series and Absolute Convergence fluids
antiderivative D 7.6: Fluid Forces
5.1: Antiderivatives and Indefinite Integration decreasing on the interval I fundamental theorem of calculus
arc length 3.3: Increasing and Decreasing Functions
5.4: The Fundamental Theorem of Calculus
7.4: Arc Length and Surface Area definite integral
9.5: Calculus and Polar Functions
11.5: The Arc Length Parameter and Curvature
5.2: The Definite Integral
G
Derivative of Inverse Functions
Arc length of a parametric curve generalized chain rule
2.7: Derivatives of Inverse Functions
9.3: Calculus and Parametric Equations 2.5: The Chain Rule
Derivative of Inverse Trigonometric
arc length parameter generalized power rule
Functions 2.5: The Chain Rule
11.5: The Arc Length Parameter and Curvature
2.7: Derivatives of Inverse Functions
Derivative of logarithmic functions H
B 2.7: Derivatives of Inverse Functions
bisection method Higher Order Derivatives
Difference Rule 2.3: Basic Differentiation Rules
1.5: Continuity
5.1: Antiderivatives and Indefinite Integration
bounds of integration. Hyperbolas
Differentiability (two variables) 9.1: Conic Sections
5.2: The Definite Integral
12.4: Differentiability and the Total Differential
hyperbolic cosine
differentials
C 4.4: Differentials
6.6: Hyperbolic Functions
Cartesian Coordinate Plane hyperbolic functions
Differentiation Rules 6.6: Hyperbolic Functions
10.1: Introduction to Cartesian Coordinates in Space 2.3: Basic Differentiation Rules
Cartesian Coordinates in Space hyperbolic sine
Directional Derivatives 6.6: Hyperbolic Functions
10.1: Introduction to Cartesian Coordinates in Space 12.6: Directional Derivatives
center of mass directrix
13.4: Center of Mass
I
9.1: Conic Sections
chain rule disk method implicit differentiation
2.5: The Chain Rule 2.6: Implicit Differentiation
7.2: Volume by Cross-Sectional Area- Disk and
4.2: Related Rates
comparison test Washer Methods
12.5: The Multivariable Chain Rule
8.3: Integral and Comparison Tests distance IMPROPER INTEGRAL
Concavity 10.1: Introduction to Cartesian Coordinates in Space
6.8: Improper Integration
3.4: Concavity and the Second Derivative Distance between a Plane and a Point increasing on the interval I
9.3: Calculus and Parametric Equations 10.6: Planes
3.3: Increasing and Decreasing Functions
Concavity (parametric equations) double integral
9.3: Calculus and Parametric Equations
independent variable
13.2: Double Integration and Volume
5.1: Antiderivatives and Indefinite Integration
CONIC SECTIONS DOUBLE INTEGRAL IN POLAR
10.1: Introduction to Cartesian Coordinates in Space
indeterminate forms
COORDINATES 6.7: L'Hopital's Rule
conics
9.1: Conic Sections
13.3: Double Integration with Polar Coordinates infinite series
8.2: Infinite Series
constant multiple rule E
5.1: Antiderivatives and Indefinite Integration
inflection point
Constant Rule eccentricity 3.4: Concavity and the Second Derivative

2.3: Basic Differentiation Rules


9.1: Conic Sections instantaneous velocity
continuity elementary function 2.1: Instantaneous Rates of Change- The Derivative
5.5: Numerical Integration 3.2: The Mean Value Theorem
1.5: Continuity
12.2: Limits and Continuity of Multivariable Ellipses Integral Test
Functions 9.1: Conic Sections 8.3: Integral and Comparison Tests
contour map extrema Integration by Parts
12.1: Introduction to Multivariable Functions 3.1: Extreme Values 6.2: Integration by Parts
convergence of a series 12.8: Extreme Values Intermediate Value Theorem
8.2: Infinite Series Extreme Value Theorem 1.5: Continuity
cross product 3.1: Extreme Values iterated integration
10.4: The Cross Product 13.1: Iterated Integrals and Area

1 https://math.libretexts.org/@go/page/45271
L orthogonal projection second derivative test
L'Hôpital's Rule 10.3: The Dot Product 3.4: Concavity and the Second Derivative
osculating circle 12.8: Extreme Values
6.7: L'Hopital's Rule
11.5: The Arc Length Parameter and Curvature second partial derivatives
lamina
12.3: Partial Derivatives
13.4: Center of Mass
P sequences
level curve of a function of two variables
8: Sequences and Series
12.1: Introduction to Multivariable Functions parabola 8.1: Sequences
level curves 9.1: Conic Sections
series
12.1: Introduction to Multivariable Functions Parallel Line 8: Sequences and Series
level surface of a function of three 10.5: Lines
Shell Method
variables Parallelogram law 7.3: The Shell Method
10.2: An Introduction to Vectors
12.1: Introduction to Multivariable Functions Simpson's rule
limacon Parametric equations 5.5: Numerical Integration
9.2: Parametric Equations
9.5: Calculus and Polar Functions skew lines
limit parametric equations of a line 10.5: Lines
10.5: Lines
1.2: Epsilon-Delta Definition of a Limit solid of revolution
limits partial derivative 7.4: Arc Length and Surface Area
12.3: Partial Derivatives 9.3: Calculus and Parametric Equations
1.1: An Introduction to Limits
1.3: Finding Limits Analytically partial fraction decomposition speed
Limits of Piecewise functions 6.5: Partial Fraction Decomposition 11.3: The Calculus of Motion
1.4: One Sided Limits partial sum squeeze theorem
local extremum 8.2: Infinite Series 1.3: Finding Limits Analytically
3.1: Extreme Values perimeter Sum Rule
Logarithmic Differentiation 4.3: Optimization 5.1: Antiderivatives and Indefinite Integration
2.6: Implicit Differentiation plane surface area
10.1: Introduction to Cartesian Coordinates in Space 7.4: Arc Length and Surface Area
M 10.6: Planes 9.3: Calculus and Parametric Equations
polar coordinates 9.5: Calculus and Polar Functions
Maclaurin series 13.5: Surface Area
9.4: Introduction to Polar Coordinates
8.8: Taylor Series surface of revolution
power rule
maxima 9.3: Calculus and Parametric Equations
2.3: Basic Differentiation Rules
4.3: Optimization 2.6: Implicit Differentiation 10.1: Introduction to Cartesian Coordinates in Space
mean value theorem 5.1: Antiderivatives and Indefinite Integration
3.2: The Mean Value Theorem power series T
midpoint rule 8.6: Power Series Tangent Line
5.3: Riemann Sums pressure 12.7: Tangent Lines, Normal Lines, and Tangent
Minima 7.6: Fluid Forces Planes
4.3: Optimization projectile motion tangent plane
12.8: Extreme Values 11.3: The Calculus of Motion 12.7: Tangent Lines, Normal Lines, and Tangent
mixed partial derivatives projection Planes
12.3: Partial Derivatives 10.3: The Dot Product tangents (parametric equations)
Multiple Integration 9.3: Calculus and Parametric Equations
13.1: Iterated Integrals and Area Q Taylor polynomials
Multivariable Chain Rule quotient rule 8.7: Taylor Polynomials
12.5: The Multivariable Chain Rule
2.4: The Product and Quotient Rules
Taylor series
Multivariable Functions 8.8: Taylor Series
12.1: Introduction to Multivariable Functions
R telescoping series
8.2: Infinite Series
N ratio test the Product Rule
8.4: Ratio and Root Tests
Newton's Method related rates
2.4: The Product and Quotient Rules
4.1: Newton's Method Total differential
4.2: Related Rates
normal lines (parametric equations) Riemann sums
12.4: Differentiability and the Total Differential
9.3: Calculus and Parametric Equations Trigonometric Integrals
5.3: Riemann Sums
normal vector Rolle’s Theorem
6.3: Trigonometric Integrals
10.6: Planes triple integral
3.2: The Mean Value Theorem
11.4: Unit Tangent and Normal Vectors 13.6: Volume Between Surfaces and Triple
root test Integration
O 8.4: Ratio and Root Tests

one sided limit U


1.4: One Sided Limits
S Unit Tangent Vector
Optimization saddle point 11.4: Unit Tangent and Normal Vectors
12.8: Extreme Values
4.3: Optimization
orthogonal decomposition Second Derivative
3.4: Concavity and the Second Derivative
10.3: The Dot Product

2 https://math.libretexts.org/@go/page/45271
V velocity W
vector 11.3: The Calculus of Motion washer method
10.2: An Introduction to Vectors Volume Between Surfaces 7.2: Volume by Cross-Sectional Area- Disk and
vector function 13.6: Volume Between Surfaces and Triple Washer Methods
Integration work
11.1: Vector–Valued Functions
Volume by Shells 7.5: Work
vectors
7.3: The Shell Method
10: Vectors

3 https://math.libretexts.org/@go/page/45271
CHAPTER OVERVIEW
7: Applications of Integration
7.1: Area Between Curves
7.2: Volume by Cross-Sectional Area- Disk and Washer Methods
7.3: The Shell Method
7.4: Arc Length and Surface Area
7.5: Work
7.6: Fluid Forces
7.E: Applications of Integration (Exercises)

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 7: Applications of Integration is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory
Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

1
7.1: Area Between Curves
We begin this chapter with a reminder of a few key concepts from Chapter 5. Let f be a continuous function on [a, b] which is
partitioned into n subintervals as
$$a<x_1 < x_2 < \cdots < x_n<x_{n+1}=b.\]
Let dx denote the length of the i
i
th
subinterval, and let c be any x-value in that subinterval. Definition 5.3.1 states that the sum
i

$$\sum_{i=1}^n f(c_i)\ dx_i\]


is a Riemann Sum. Riemann Sums are often used to approximate some quantity (area, volume, work, pressure, etc.). The
approximation becomes exact by taking the limit
$$\lim_{||dx_i||\to0} \sum_{i=1}^n f(c_i) \ dx_i,\]
where || dx i || the length of the largest subinterval in the partition. Theorem 5.3.2 connects limits of Riemann Sums to definite
integrals:
$$\lim_{||dx_i||\to0} \sum_{i=1}^n f(c_i)\ dx_i = \int_a^b f(x)\ dx.\]
Finally, the Fundamental Theorem of Calculus states how definite integrals can be evaluated using antiderivatives.
This chapter employs the following technique to a variety of applications. Suppose the value Q of a quantity is to be calculated. We
first approximate the value of Q using a Riemann Sum, then find the exact value via a definite integral. We spell out this technique
in the following Key Idea.

Key Idea 22: Definite Integral Strategy


Let a quantity be given whose value Q is to be computed.
1. Divide the quantity into n smaller "subquantities" of value Q . i

2. Identify a variable x and function f (x) such that each subquantity can be approximated with the product f (c ) dx , wherei i

dx represents a small change in x. Thus Q ≈ f (c ) dx . A sample approximation f (c ) dx of Q is called a differential


i i i i i i i

element.
3. Recognize that Q = ∑ Q ≈ ∑ f (c ) dx , which is a Riemann Sum.
n

i=1 i
n

i=1 i i

b
4. Taking the appropriate limit gives Q = ∫ a
f (x) dx

This Key Idea will make more sense after we have had a chance to use it several times. We begin with Area Between Curves,
which we addressed briefly in Section 5.5.4.

Area Between Curves


We are often interested in knowing the area of a region. Forget momentarily that we addressed this already in Section 5.5.4 and
approach it instead using the technique described in Key Idea 22.
Let Q be the area of a region bounded by continuous functions f and g . If we break the region into many subregions, we have an
obvious equation:
Total Area= sum of the areas of the subregions. (7.1.1)

The issue to address next is how to systematically break a region into subregions. A graph will help. Consider Figure 7.1.1a where
a region between two curves is shaded. While there are many ways to break this into subregions, one particularly efficient way is to
"slice" it vertically, as shown in Figure 7.1.1b, into n equally spaced slices.

7.1.1 https://math.libretexts.org/@go/page/4192
Figure 7.1.1 : Subdividing a region into vertical slices and approximating the areas with rectangles.
We now approximate the area of a slice. Again, we have many options, but using a rectangle seems simplest. Picking any x-value
c in the i slice, we set the height of the rectangle to be f (c ) − g(c ) , the difference of the corresponding y -values. The width of
th
i i i

the rectangle is a small difference in x-values, which we represent with dx. Figure 7.1.1c shows sample points c chosen in each i

subinterval and appropriate rectangles drawn. (Each of these rectangles represents a differential element.) Each slice has an area
approximately equal to (f (c ) − g(c )) dx ; hence, the total area is approximately the Riemann Sum
i i

$$Q = \sum_{i=1}^n \big(f(c_i)-g(c_i)\big)\ dx.\]


b
Taking the limit as n → ∞ gives the exact area as ∫ a
(f (x) − g(x)) dx.

Theroem 7.1.1: Area Between Curves

Let f (x) and g(x) be continuous functions defined on [a, b] where f (x) ≥ g(x) for all x in . The area of the region
[a, b]

bounded by the curves y = f (x), y = g(x) and the lines x = a and x = b is


$$\int_a^b \big(f(x)-g(x)\big)\ dx.\]

Example 7.1.1: Finding area enclosed by curves


1
Find the area of the region bounded by f (x) = sin x + 2 , g(x) = cos(2x) − 1 , x =0 and x = 4π , as shown in Figure
2
7.1.2.

Figure 7.1.2 : Graphing an enclosed region in Example 7.1.1.


Solution
The graph verifies that the upper boundary of the region is given by f and the lower bound is given by g . Therefore the area of
the region is the value of the integral

7.1.2 https://math.libretexts.org/@go/page/4192
4π 4π
1
∫ (f (x) − g(x)) dx =∫ ( sin x + 2 − ( cos(2x) − 1)) dx
0 0 2

1 4π

= − cos x − sin(2x) + 3x

4 0

2
= 12π ≈ 37.7 units .

Example 7.1.2: Finding total area enclosed by curves

Find the total area of the region enclosed by the functions f (x) = −2x + 5 and g(x) = x
3
− 7x
2
+ 12x − 3 as shown in
Figure 7.1.3.

FIgure 7.1.3 : Graphing a region enclosed by two functions in Example 7.1.2.


Solution
4
A quick calculation shows that f = g at x = 1, 2 and 4. One can proceed thoughtlessly by computing ∫ (f (x) − g(x)) dx , 1

but this ignores the fact that on [1, 2], g(x) > f (x) . (In fact, the thoughtless integration returns −9/4, hardly the expected
value of an area.) Thus we compute the total area by breaking the interval [1, 4] into two subintervals, [1, 2] and [2, 4] and
using the proper integrand in each.
2 4

Total Area =∫ (g(x) − f (x)) dx + ∫ (f (x) − g(x)) dx


1 2

2 4
3 2 3 2
=∫ (x − 7x + 14x − 8) dx + ∫ (−x + 7x − 14x + 8) dx
1 2

= 5/12 + 8/3

2
= 37/12 = 3.083 units .

The previous example makes note that we are expecting area to be positive. When first learning about the definite integral, we
interpreted it as "signed area under the curve," allowing for "negative area." That doesn't apply here; area is to be positive.
The previous example also demonstrates that we often have to break a given region into subregions before applying Theorem
7.1.1. The following example shows another situation where this is applicable, along with an alternate view of applying the

Theorem.

Example 7.1.3: Finding area: integrating with respect to y

Find the area of the region enclosed by the functions y = √−


x + 2 , y = −(x − 1 )
2
+3 and y = 2 , as shown in Figure 7.1.4.

7.1.3 https://math.libretexts.org/@go/page/4192
Figure 7.1.4 : Graphing a region for Example 7.1.3.
Solution
We give two approaches to this problem. In the first approach, we notice that the region's "top" is defined by two different
curves. On [0, 1], the top function is y = √−
x + 2 ; on [1, 2], the top function is y = −(x − 1 ) + 3 .
2

Thus we compute the area as the sum of two integrals:


1 2
− 2
Total Area =∫ ((√x + 2) − 2) dx + ∫ (( − (x − 1 ) + 3) − 2) dx
0 1

= 2/3 + 2/3

= 4/3.

The second approach is clever and very useful in certain situations. We are used to viewing curves as functions of x; we input
an x-value and a y -value is returned. Some curves can also be described as functions of y : input a y -value and an x-value is
returned. We can rewrite the equations describing the boundary by solving for x:
$$y=\sqrt{x}+2 \quad \Rightarrow\quad x=(y-2)^2\]
$$y=-(x-1)^2+3 \quad \Rightarrow \quad x=\sqrt{3-y}+1.\]

Figure 7.1.5 :The region used in Example 7.1.3 with boundaries relabeled as functions of y .
Figure 7.1.5 shows the region with the boundaries relabeled. A differential element, a horizontal rectangle, is also pictured.
The width of the rectangle is a small change in y : Δy. The height of the rectangle is a difference in x-values. The "top" x-
value is the largest value, i.e., the rightmost. The "bottom" x-value is the smaller, i.e., the leftmost. Therefore the height of the
rectangle is
$$\big(\sqrt{3-y}+1\big) - (y-2)^2.\]

7.1.4 https://math.libretexts.org/@go/page/4192
The area is found by integrating the above function with respect to y with the appropriate bounds. We determine these by
considering the y -values the region occupies. It is bounded below by y = 2 , and bounded above by y = 3 . That is, both the
"top" and "bottom" functions exist on the y interval [2, 3]. Thus
3
− −−− 2
Total Area =∫ (√ 3 − y + 1 − (y − 2 ) ) dy
2

2 1 3
3/2 3 ∣
=(− (3 − y ) +y − (y − 2 ) )
∣2
3 3

= 4/3.

This calculus--based technique of finding area can be useful even with shapes that we normally think of as "easy." Example 7.1.4
1
computes the area of a triangle. While the formula " × base × height " is well known, in arbitrary triangles it can be nontrivial to
2
compute the height. Calculus makes the problem simple.

Example 7.1.4: Finding the area of a triangle

Compute the area of the regions bounded by the lines


1 5
y = x +1 , y = −2x + 7 and y = − x+ , as shown in Figure 7.1.6.
2 2

Figure 7.1.6 : Graphing a triangular region in Example 7.1.4


Solution
Recognize that there are two "top" functions to this region, causing us to use two definite integrals.
2 3
1 5 1 5
Total Area =∫ ((x + 1) − (− x+ )) dx + ∫ ((−2x + 7) − (− x+ )) dx
1
2 2 2
2 2

= 3/4 + 3/4

= 3/2.

We can also approach this by converting each function into a function of y . This also requires 2 integrals, so there isn't really
any advantage to doing so. We do it here for demonstration purposes.
7 −y
The "top" function is always x = while there are two "bottom" functions. Being mindful of the proper integration
2
bounds, we have

7.1.5 https://math.libretexts.org/@go/page/4192
2 3
7 −y 7 −y
Total Area =∫ ( − (5 − 2y)) dy + ∫ ( − (y − 1)) dy
1 2 2 2

= 3/4 + 3/4

= 3/2.

Of course, the final answer is the same. (It is interesting to note that the area of all 4 subregions used is 3/4. This is
coincidental.)

While we have focused on producing exact answers, we are also able to make approximations using the principle of Theorem
7.1.1. The integrand in the theorem is a distance ("top minus bottom"); integrating this distance function gives an area. By taking

discrete measurements of distance, we can approximate an area using numerical integration techniques developed in Section ??? .
The following example demonstrates this.

Example 7.1.5: Numerically approximating area

To approximate the area of a lake, shown in Figure 7.1.7a, the "length" of the lake is measured at 200-foot increments as
shown in Figure 7.1.7b, where the lengths are given in hundreds of feet. Approximate the area of the lake.
Solution
The measurements of length can be viewed as measuring "top minus bottom" of two functions. The exact answer is found by
12
integrating ∫ (f (x) − g(x)) dx , but of course we don't know the functions f and g . Our discrete measurements instead
0

allow us to approximate.

Figure 7.1.7 : (a) A sketch of a lake, and (b) the lake with length measurements.
We have the following data points:
$$(0,0),\ (2,2.25),\ (4,5.08),\ (6,6.35),\ (8,5.21),\ (10,2.76),\ (12,0).\]
b −a
We also have that dx = =2 , so Simpson's Rule gives
n

2
Area ≈ (1 ⋅ 0 + 4 ⋅ 2.25 + 2 ⋅ 5.08 + 4 ⋅ 6.35 + 2 ⋅ 5.21 + 4 ⋅ 2.76 + 1 ⋅ 0)
3

¯
¯¯ 2
= 44.01 3 units .

Since the measurements are in hundreds of feet, units = (100 ft) = 10, 000 ft , giving a total area of 440, 133 ft . (Since
2 2 2 2

we are approximating, we'd likely say the area was about 440, 000 ft , which is a little more than 10 acres.)
2

In the next section we apply our applications--of--integration techniques to finding the volumes of certain solids.

7.1.6 https://math.libretexts.org/@go/page/4192
Contributors and Attributions
Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/
Integrated by Justin Marshall.

This page titled 7.1: Area Between Curves is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory
Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

7.1.7 https://math.libretexts.org/@go/page/4192
7.2: Volume by Cross-Sectional Area- Disk and Washer Methods
The volume of a general right cylinder, as shown in Figure 7.2.1, is
Area of the base × height. (7.2.1)

We can use this fact as the building block in finding volumes of a variety of shapes.
Given an arbitrary solid, we can approximate its volume by cutting it into n thin slices. When the slices are thin, each slice can be
approximated well by a general right cylinder. Thus the volume of each slice is approximately its cross-sectional area × thickness.
(These slices are the differential elements.)

Figure 7.2.1 : The volume of a general right cylinder.


By orienting a solid along the x-axis, we can let A(x ) represent the cross-sectional area
i

of the i slice, and let dx represent the thickness of this slice (the thickness is a small change in x). The total volume of the solid
th
i

is approximately:
n

Volume ≈ ∑ [Area × thickness] (7.2.2)

i=1

= ∑ A(xi ) dxi . (7.2.3)

i=1

Recognize that this is a Riemann Sum. By taking a limit (as the thickness of the slices goes to 0) we can find the volume exactly.

Theorem 7.2.1: Volume By Cross-Sectional Area

The volume V of a solid, oriented along the x-axis with cross-sectional area A(x) from x = a to x = b , is
$$V = \int_a^b A(x)\ dx.\]

Example 7.2.1: Finding the volume of a solid

Find the volume of a pyramid with a square base of side length 10 in and a height of 5 in.
Solution
There are many ways to "orient" the pyramid along the x-axis; Figure 7.2.2 gives one such way, with the pointed top of the
pyramid at the origin and the x-axis going through the center of the base.

7.2.1 https://math.libretexts.org/@go/page/4193
Figure 7.2.2 : Orienting a pyramid along the x-axis in Example 7.2.1.
Each cross section of the pyramid is a square; this is a sample differential element. To determine its area A(x) , we need to
determine the side lengths of the square.
When x = 5 , the square has side length 10; when x = 0 , the square has side length 0. Since the edges of the pyramid are lines,
it is easy to figure that each cross-sectional square has side length 2x, giving A(x) = (2x ) = 4x .2 2

If one were to cut a slice out of the pyramid at x = 3 , as shown in Figure 7.2.3, one would have a shape with square bottom
and top with sloped sides. If the slice were thin, both the bottom and top squares would have sides lengths of about 6, and thus
the cross--sectional area of the bottom and top would be about 36in . Letting Δx represent the thickness of the slice, the
2
i

volume of this slice would then be about 36Δx \(in^3\).


i

Figure 7.2.3 : Cutting a slice in they pyramid in Example 7.2.1 at x = 3 .


Cutting the pyramid into n slices divides the total volume into n equally--spaced smaller pieces, each with volume (2x ) Δx, i
2

where x is the approximate location of the slice along the x-axis and Δx represents the thickness of each slice. One can
i

approximate total volume of the pyramid by summing up the volumes of these slices:
$$\text{Approximate volume } = \sum_{i=1}^n (2x_i)^2\Delta x.\]
Taking the limit as n → ∞ gives the actual volume of the pyramid; recoginizing this sum as a Riemann Sum allows us to find
the exact answer using a definite integral, matching the definite integral given by Theorem 7.2.1.
We have
n

2
V = lim ∑(2 xi ) Δx (7.2.4)
n→∞
i=1

5
2
=∫ 4x dx (7.2.5)
0

4 5
3∣
= x (7.2.6)

3 0

500
3 3
= in ≈ 166.67 in . (7.2.7)
3

7.2.2 https://math.libretexts.org/@go/page/4193
We can check our work by consulting the general equation for the volume of a pyramid (see the back cover under "Volume of
A General Cone"):
1
× area of base × height. (7.2.8)
3

Certainly, using this formula from geometry is faster than our new method, but the calculus--based method can be applied to
much more than just cones.

An important special case of Theorem 7.2.1 is when the solid is a solid of revolution, that is, when the solid is formed by rotating
a shape around an axis.
Start with a function y = f (x) from x = a to x = b . Revolving this curve about a horizontal axis creates a three-dimensional solid
whose cross sections are disks (thin circles). Let R(x) represent the radius of the cross-sectional disk at x; the area of this disk is
πR(x) . Applying Theorem 7.2.1 gives the Disk Method.
2

Key Idea 23: The Disk Method

Let a solid be formed by revolving the curve y = f (x) from x = a to x =b around a horizontal axis, and let R(x) be the
radius of the cross-sectional disk at x. The volume of the solid is
$$V = \pi \int_a^b R(x)^2\ dx.\]

Example 7.2.2: Finding volume using the Disk Method

Find the volume of the solid formed by revolving the curve y = 1/x, from x = 1 to x = 2 , around the x-axis.
Solution
A sketch can help us understand this problem. In Figure 7.2.4a the curve y = 1/x is sketched along with the differential
element -- a disk -- at x with radius R(x) = 1/x. In Figure 7.2.4b the whole solid is pictured, along with the differential
element.
The volume of the differential element shown in part (a) of the figure is approximately πR(x ) Δx, where R(x ) is the radius
i
2
i

of the disk shown and Δx is the thickness of that slice. The radius R(x ) is the distance from the x-axis to the curve, hence
i

R(x ) = 1/ x .
i i

Figure 7.2.4 : Sketching a solid in Example 7.2.2.


Slicing the solid into n equally--spaced slices, we can approximate the total volume by adding up the approximate volume of
each slice:
$$\text{Approximate volume } = \sum_{i=1}^n \pi \left(\frac1{x_i}\right)^2\Delta x.\]
Taking the limit of the above sum as n → ∞ gives the actual volume; recognizing this sum as a Riemann sum allows us to
evaluate the limit with a definite integral, which matches the formula given in Key Idea 23:

7.2.3 https://math.libretexts.org/@go/page/4193
n 2
1
V = lim ∑ π ( ) Δx (7.2.9)
n→∞ xi
i=1

2 2
1
=π∫ ( ) dx (7.2.10)
1 x
2
1
=π∫ dx (7.2.11)
2
1 x

1 2

= π [− ] (7.2.12)

x 1

1
= π [− − (−1)] (7.2.13)
2
π
3
= units . (7.2.14)
2

While Key Idea 23 is given in terms of functions of x, the principle involved can be applied to functions of y when the axis of
rotation is vertical, not horizontal. We demonstrate this in the next example.

Example 7.2.3: Finding volume using the Disk Method

Find the volume of the solid formed by revolving the curve y = 1/x, from x = 1 to x = 2 , about the y -axis.
Solution
Since the axis of rotation is vertical, we need to convert the function into a function of y and convert the x-bounds to y -
bounds. Since y = 1/x defines the curve, we rewrite it as x = 1/y. The bound x = 1 corresponds to the y -bound y = 1 , and
the bound x = 2 corresponds to the y -bound y = 1/2.
Thus we are rotating the curve x = 1/y, from y = 1/2 to y = 1 about the y -axis to form a solid. The curve and sample
differential element are sketched in Figure 7.2.5a, with a full sketch of the solid in Figure 7.2.5b.

Figure 7.2.5 : Sketching a solid in Example 7.2.3.


We integrate to find the volume:
1
1
V =π∫ dy (7.2.15)
2
1/2 y

π 1

=− (7.2.16)
∣1/2
y
3
= π units . (7.2.17)

We can also compute the volume of solids of revolution that have a hole in the center. The general principle is simple: compute the
volume of the solid irrespective of the hole, then subtract the volume of the hole. If the outside radius of the solid is R(x) and the
inside radius (defining the hole) is r(x), then the volume is

7.2.4 https://math.libretexts.org/@go/page/4193
$$V = \pi\int_a^b R(x)^2 \ dx - \pi\int_a^b r(x)^2\ dx = \pi\int_a^b \left(R(x)^2-r(x)^2\right)\ dx.\]

Figure 7.2.6 : Establishing the Washer Method; see also Figure 7.2.7.
One can generate a solid of revolution with a hole in the middle by revolving a region about an axis. Consider Figure 7.2.6a, where
a region is sketched along with a dashed, horizontal axis of rotation. By rotating the region about the axis, a solid is formed as
sketched in Figure 7.2.6b. The outside of the solid has radius R(x), whereas the inside has radius r(x). Each cross section of this
solid will be a washer (a disk with a hole in the center) as sketched in Figure 7.2.6b. This leads us to the Washer Method.

Figure 7.2.7 : Establishing the Washer Method; see also Figure 7.2.6.

Key Idea 24: The Washer Method

Let a region bounded by y = f (x), y = g(x) , x = a and x = b be rotated about a horizontal axis that does not intersect the
region, forming a solid. Each cross section at x will be a washer with outside radius R(x) and inside radius r(x). The volume
of the solid is
b
2 2
V =π∫ (R(x ) − r(x ) ) dx. (7.2.18)
a

Even though we introduced it first, the Disk Method is just a special case of the Washer Method with an inside radius of r(x) = 0 .

Example 7.2.4: Finding volume with the Washer Method

Find the volume of the solid formed by rotating the region bounded by y = x 2
− 2x + 2 and y = 2x − 1 about the x-axis.
Solution
A sketch of the region will help, as given in Figure 7.2.8a.

7.2.5 https://math.libretexts.org/@go/page/4193
Figure 7.2.8 : Sketching the differential element and solid in Example 7.2.4.
Rotating about the x-axis will produce cross sections in the shape of washers, as shown in Figure 7.2.8b; the complete solid is
shown in part (c). The outside radius of this washer is R(x) = 2x + 1 ; the inside radius is r(x) = x − 2x + 2 . As the region
2

is bounded from x = 1 to x = 3 , we integrate as follows to compute the volume.


3
2 2 2
V =π∫ ((2x − 1 ) − (x − 2x + 2 ) ) dx (7.2.19)
1

3
4 3 2
=π∫ (−x + 4x − 4x + 4x − 3) dx (7.2.20)
1

1 4 3
5 4 3 2 ∣
= π[ − x +x − x + 2x − 3x] (7.2.21)

5 3 1

104
3
= π ≈ 21.78 units . (7.2.22)
15

When rotating about a vertical axis, the outside and inside radius functions must be functions of y .

Example 7.2.5: Finding volume with the Washer Method

Find the volume of the solid formed by rotating the triangular region with vertices at (1, 1), (2, 1) and (2, 3) about the y -axis.
Solution
The triangular region is sketched in Figure 7.2.9; the differential element is sketched in (b) and the full solid is drawn in (c).
They help us establish the outside and inside radii. Since the axis of rotation is vertical, each radius is a function of y .
The outside radius R(y) is formed by the line connecting (2, 1) and (2, 3); it is a constant function, as regardless of the y -
value the distance from the line to the axis of rotation is 2. Thus R(y) = 2 .

Figure 7.2.9 : Sketching the solid in Example 7.2.5.

7.2.6 https://math.libretexts.org/@go/page/4193
The inside radius is formed by the line connecting (1, 1) and (2, 3). The equation of this line is y = 2x − 1 , but we need to
refer to it as a function of y . Solving for x gives r(y) = (y + 1) .
1

We integrate over the y -bounds of y = 1 to y = 3 . Thus the volume is


3
1 2
2
V =π∫ (2 −( (y + 1)) ) dy (7.2.23)
1
2
3
1 2
1 15
=π∫ (− y − y+ ) dy (7.2.24)
1
4 2 4

1 1 15 3
3 2 ∣
= π[ − y − y + y] (7.2.25)

12 4 4 1

10 3
= π ≈ 10.47 units . (7.2.26)
3

This section introduced a new application of the definite integral. Our default view of the definite integral is that it gives "the area
under the curve." However, we can establish definite integrals that represent other quantities; in this section, we computed volume.
The ultimate goal of this section is not to compute volumes of solids. That can be useful, but what is more useful is the
understanding of this basic principle of integral calculus, outlined in Key Idea 24: to find the exact value of some quantity,
we start with an approximation (in this section, slice the solid and approximate the volume of each slice),
then make the approximation better by refining our original approximation (i.e., use more slices),
then use limits to establish a definite integral which gives the exact value.
We practice this principle in the next section where we find volumes by slicing solids in a different way.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/
Integrated by Justin Marshall.

This page titled 7.2: Volume by Cross-Sectional Area- Disk and Washer Methods is shared under a CC BY-NC 3.0 license and was authored,
remixed, and/or curated by Gregory Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a
detailed edit history is available upon request.

7.2.7 https://math.libretexts.org/@go/page/4193
7.3: The Shell Method
Often a given problem can be solved in more than one way. A particular method may be chosen out of convenience, personal
preference, or perhaps necessity. Ultimately, it is good to have options.
The previous section introduced the Disk and Washer Methods, which computed the volume of solids of revolution by integrating
the cross--sectional area of the solid. This section develops another method of computing volume, the Shell Method. Instead of
slicing the solid perpendicular to the axis of rotation creating cross-sections, we now slice it parallel to the axis of rotation, creating
"shells."
Consider Figure 7.3.1, where the region shown in (a) is rotated around the y -axis forming the solid shown in (b). A small slice of
the region is drawn in (a), parallel to the axis of rotation. When the region is rotated, this thin slice forms a cylindrical shell, as
pictured in part (c) of the figure. The previous section approximated a solid with lots of thin disks (or washers); we now
approximate a solid with many thin cylindrical shells.

Figure 7.3.1 : Introducing the Shell Method.

7.3.1 https://math.libretexts.org/@go/page/4194
Figure 7.3.1 (d): A dynamic version of this figure created using CalcPlot3D.
To compute the volume of one shell, first consider the paper label on a soup can with radius r and height h . What is the area of this
label? A simple way of determining this is to cut the label and lay it out flat, forming a rectangle with height h and length 2πr.
Thus the area is A = 2πrh ; see Figure 7.3.2a.
Do a similar process with a cylindrical shell, with height h , thickness Δx, and approximate radius r. Cutting the shell and laying it
flat forms a rectangular solid with length 2πr, height h and depth dx. Thus the volume is V ≈ 2πrh dx ; see Figure 7.3.2c. (We
say "approximately" since our radius was an approximation.)
By breaking the solid into n cylindrical shells, we can approximate the volume of the solid as
$$V = \sum_{i=1}^n 2\pi r_ih_i\ dx_i,\]
where r , h and dx are the radius, height and thickness of the i
i i i
th
shell, respectively.
This is a Riemann Sum. Taking a limit as the thickness of the shells approaches 0 leads to a definite integral.

7.3.2 https://math.libretexts.org/@go/page/4194
Figure 7.3.2 : Determining the volume of a thin cylindrical shell.}\label{fig:soupcan}

Key Idea 25: Shell Method


Let a solid be formed by revolving a region R , bounded by x = a and x = b , around a vertical axis. Let r(x) represent the
distance from the axis of rotation to x (i.e., the radius of a sample shell) and let h(x) represent the height of the solid at x (i.e.,
the height of the shell). The volume of the solid is
b

V = 2π ∫ r(x)h(x) dx. (7.3.1)


a

Special Cases:
1. When the region R is bounded above by y = f (x) and below by y = g(x) , then h(x) = f (x) − g(x) .
2. When the axis of rotation is the y -axis (i.e., x = 0 ) then r(x) = x .
Let's practice using the Shell Method.

Example 7.3.1: Finding volume using the Shell Method

Find the volume of the solid formed by rotating the region bounded by y = 0 , y = 1/(1 + x 2
) , x = 0 and x = 1 about the y -
axis.
Solution
This is the region used to introduce the Shell Method in Figure 7.3.1, but is sketched again in Figure 7.3.3 for closer reference.
A line is drawn in the region parallel to the axis of rotation representing a shell that will be carved out as the region is rotated
about the y -axis. (This is the differential element.)

Figure 7.3.3 : Graphing a region in Example 7.3.1.


The distance this line is from the axis of rotation determines r(x); as the distance from x to the y -axis is x, we have r(x) = x .
The height of this line determines h(x); the top of the line is at y = 1/(1 + x ) , whereas the bottom of the line is at y = 0 .
2

Thus h(x) = 1/(1 + x ) − 0 = 1/(1 + x ) . The region is bounded from x = 0 to x = 1 , so the volume is
2 2

7.3.3 https://math.libretexts.org/@go/page/4194
1
x
V = 2π ∫ dx. (7.3.2)
2
0 1 +x

This requires substitution. Let u = 1 +x


2
, so du = 2x dx . We also change the bounds: u(0) = 1 and u(1) = 2 . Thus we
have:
2
1
=π∫ du
1
u

2

= π ln u

1

= π ln 2 − π ln 1

3
= π ln 2 ≈ 2.178 units .

Note: in order to find this volume using the Disk Method, two integrals would be needed to account for the regions above and
below y = 1/2.

With the Shell Method, nothing special needs to be accounted for to compute the volume of a solid that has a hole in the middle, as
demonstrated next.

Example 7.3.2: Finding volume using the Shell Method

Find the volume of the solid formed by rotating the triangular region determined by the points ,
(0, 1) (1, 1) and (1, 3) about
the line x = 3 .
Solution
The region is sketched in Figure 7.3.4a along with the differential element, a line within the region parallel to the axis of
rotation. In part (b) of the figure, we see the shell traced out by the differential element, and in part (c) the whole solid is
shown.

Figure 7.3.4 : Graphing a region in Example 7.3.2


The height of the differential element is the distance from y = 1 to y = 2x + 1 , the line that connects the points (0, 1) and
(1, 3). Thus h(x) = 2x + 1 − 1 = 2x . The radius of the shell formed by the differential element is the distance from x to

x = 3 ; that is, it is r(x) = 3 − x . The x-bounds of the region are x = 0 to x = 1 , giving

V = 2π ∫ (3 − x)(2x) dx
0

1
2
= 2π ∫ (6x − 2 x ) dx
0

2 1
2 3 ∣
= 2π (3 x − x )

3 0

14
3
= π ≈ 14.66 units .
3

7.3.4 https://math.libretexts.org/@go/page/4194
When revolving a region around a horizontal axis, we must consider the radius and height functions in terms of y , not x.

Example 7.3.3: Finding volume using the Shell Method

Find the volume of the solid formed by rotating the region given in Example 7.3.2 about the x-axis.
Solution
The region is sketched in Figure 7.3.5a with a sample differential element. In part (b) of the figure the shell formed by the
differential element is drawn, and the solid is sketched in (c). (Note that the triangular region looks "short and wide" here,
whereas in the previous example the same region looked "tall and narrow." This is because the bounds on the graphs are
different.)
1 1
The height of the differential element is an x -distance, between x = y− and x =1 . Thus
2 2
1 1 1 3
h(y) = 1 − ( y− ) =− y+ . The radius is the distance from y to the x-axis, so r(y) = y . The y bounds of the region
2 2 2 2
are y = 1 and y = 3 , leading to the integral
3
1 3
V = 2π ∫ [y (− y+ )] dy
1
2 2

3
1 2
3
= 2π ∫ [− y + y] dy
1
2 2

1 3 3
3 2 ∣
= 2π [− y + y ]

6 4 1

9 7
= 2π [ − ]
4 12

10
3
= π ≈ 10.472 units .
3

Figure 7.3.5 : Graphing a region in Example 7.3.3

At the beginning of this section it was stated that "it is good to have options." The next example finds the volume of a solid rather
easily with the Shell Method, but using the Washer Method would be quite a chore.

Example 7.3.4: Finding volume using the Shell Method

Find the volume of the solid formed by revolving the region bounded by y = sin x and the x-axis from x = 0 to x = π about
the y -axis.
Solution
The region and a differential element, the shell formed by this differential element, and the resulting solid are given in Figure
7.3.6.

7.3.5 https://math.libretexts.org/@go/page/4194
Figure 7.3.6 : Graphing a region in Example 7.3.4
The radius of a sample shell is r(x) = x ; the height of a sample shell is h(x) = sin x , each from x =0 to x =π . Thus the
volume of the solid is
π

V = 2π ∫ x sin x dx. (7.3.3)


0

This requires Integration By Parts. Set u = x and dv = sin x dx ; we leave it to the reader to fill in the rest. We have:
π
π

= 2π[ − x cos x +∫ cos x dx]

0
0

π

= 2π[π + sin x ]

0

= 2π[π + 0]

2 3
= 2π ≈ 19.74 units .

Note that in order to use the Washer Method, we would need to solve y = sin x for x, requiring the use of the arcsine function.
We leave it to the reader to verify that the outside radius function is R(y) = π − arcsin y and the inside radius function is
r(y) = arcsin y . Thus the volume can be computed as

$$\pi\int_0^1 \Big[ (\pi-\arcsin y)^2-(\arcsin y)^2\Big]\ dy.\]


This integral isn't terrible given that the arcsin
2
y terms cancel, but it is more onerous than the integral created by the Shell
Method.

We end this section with a table summarizing the usage of the Washer and Shell Methods.

Key Idea 26: Summary of the Washer and Shell Methods


Let a region R be given with x-bounds x = a and x = b and y -bounds y = c and y = d .
Washer Method Shell Method

b d
2 2
Horizontal Axis π∫ (R(x ) − r(x ) ) dx 2π ∫ r(y)h(y) dy
a c

d b
2 2
Vertical Axis π∫ (R(y ) − r(y ) ) dy 2π ∫ r(x)h(x) dx
c a

As in the previous section, the real goal of this section is not to be able to compute volumes of certain solids. Rather, it is to be able
to solve a problem by first approximating, then using limits to refine the approximation to give the exact value. In this section, we
approximate the volume of a solid by cutting it into thin cylindrical shells. By summing up the volumes of each shell, we get an
approximation of the volume. By taking a limit as the number of equally spaced shells goes to infinity, our summation can be
evaluated as a definite integral, giving the exact value.

7.3.6 https://math.libretexts.org/@go/page/4194
We use this same principle again in the next section, where we find the length of curves in the plane.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/
Integrated by Justin Marshall.

This page titled 7.3: The Shell Method is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman
et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

7.3.7 https://math.libretexts.org/@go/page/4194
7.4: Arc Length and Surface Area
Arc Length
In previous sections we have used integration to answer the following questions:
1. Given a region, what is its area?
2. Given a solid, what is its volume?
In this section, we address a related question: Given a curve, what is its length? This is often referred to as arc length.
Consider the graph of y = sin x on [0, π] given in Figure 7.4.1a. How long is this curve? That is, if we were to use a piece of string
to exactly match the shape of this curve, how long would the string be?
As we have done in the past, we start by approximating; later, we will refine our answer using limits to get an exact solution.
The length of straight--line segments is easy to compute using the Distance Formula. We can approximate the length of the given
curve by approximating the curve with straight lines and measuring their lengths.

Figure 7.4.1 : Graphing y = sin x on [0, π] and approximating the curve with line segments.
In Figure 7.4.1b, the curve y = sin x has been approximated with 4 line segments (the interval [0, π] has been divided into 4
equally--lengthed subintervals). It is clear that these four line segments approximate y = sin x very well on the first and last
subinterval, though not so well in the middle. Regardless, the sum of the lengths of the line segments is 3.79, so we approximate
the arc length of y = sin x on [0, π] to be 3.79.
In general, we can approximate the arc length of y = f (x) on [a, b] in the following manner. Let
a = x1 < x2 < … < xn < xn+1 = b be a partition of [a, b] into n subintervals. Let dx represent the length of the i
i
th

subinterval [x i, .
xi+1 ]

Figure 7.4.2 : Zooming in on the i th


subinterval [x
i, xi+1 ] of a partition of [a, b].
Figure 7.4.2 zooms in on the i subinterval where y = f (x) is approximated by a straight line segment. The dashed lines show
th

that we can view this line segment as they hypotenuse of a right triangle whose sides have length dx and dy . Using the i i

7.4.1 https://math.libretexts.org/@go/page/4195
−−−−−−−−−
Pythagorean Theorem, the length of this line segment is √dx
2
i
+ Δy
2
i
. Summing over all subintervals gives an arc length
approximation
n
−−−−−−−−−
2 2
L ≈ ∑ √ dx + Δy . (7.4.1)
i i

i=1

As shown here, this is not a Riemann Sum. While we could conclude that taking a limit as the subinterval length goes to zero gives
the exact arc length, we would not be able to compute the answer with a definite integral. We need first to do a little algebra.
In the above expression factor out a dx term: 2
i

− −−−−−−−−−−− −

n n 2
−−−−−−−−−  Δy
2 2 2 i
∑ √ dx + Δy = ∑  dx (1 + ). (7.4.2)
i i i 2
⎷ dx
i=1 i=1 i

Now pull the dx term out of the square root:


2
i

−−−−−−−
n 2
Δy
i
= ∑ √1 + dxi . (7.4.3)
2
dx
i=1 i

This is nearly a Riemann Sum. Consider the Δy /dx term. The expression Δy /dx measures the "change in y /change in x," that
2
i
2
i i i

is, the "rise over run" of f on the i subinterval. The Mean Value Theorem of Differentiation (Theorem 3.2.1) states that there is a
th

c in the i subinterval where f (c ) = Δy /dx . Thus we can rewrite our above expression as:
th ′
i i i i

n
−−−−−−−−−
′ 2
= ∑ √ 1 + f (ci ) dxi . (7.4.4)

i=1

This is a Riemann Sum. As long as f is continuous, we can invoke Theorem 5.3.2 and conclude

b −−−−−−−−
′ 2
=∫ √ 1 + f (x ) dx. (7.4.5)
a

Key Idea 27: Arc Length


Let f be differentiable on an open interval containing [a, b] , where f

is also continuous on . Then the arc length of
[a, b] f

from x = a to x = b is
b −−−−−−−−
′ 2
L =∫ √ 1 + f (x ) dx. (7.4.6)
a

As the integrand contains a square root, it is often difficult to use the formula in Key Idea 27 to find the length exactly. When exact
answers are difficult to come by, we resort to using numerical methods of approximating definite integrals. The following examples
will demonstrate this.

Example 7.4.1: Finding arc length

Find the arc length of f (x) = x 3/2


from x = 0 to x = 4 .

7.4.2 https://math.libretexts.org/@go/page/4195
Figure 7.4.3 : A graph of f (x) = x 3/2
from Example 7.4.1
Solution
We begin by finding f ′
(x) =
3

2
1/2
x . Using the formula, we find the arc length L as
−−−−−−−−−−−
4 2
3
1/2
L =∫ √1 + ( x ) dx (7.4.7)
0
2

4 −−−−− −
9
=∫ √1 + x dx (7.4.8)
0
4

4 1/2
9
=∫ (1 + x) dx (7.4.9)
0 4

3/2
2 4 9 4

= (1 + x) (7.4.10)

3 9 4 0

8
3/2
= (10 − 1) ≈ 9.07units. (7.4.11)
27

Example 7.4.2: Finding arc length

Find the arc length of f (x) = 1

8
x
2
− ln x from x = 1 to x = 2 .

Figure 7.4.4 : A graph of f (x) = 1

8
x
2
− ln x from Example 7.4.2.
Solution
This function was chosen specifically because the resulting integral can be evaluated exactly. We begin by finding
f (x) = x/4 − 1/x . The arc length is

7.4.3 https://math.libretexts.org/@go/page/4195
−−−−−−−−−−−−
2 2
x 1
L =∫ √1 + ( − ) dx (7.4.12)
1
4 x

−−−−−−−−−−−−−−
2
x2 1 1
=∫ √ 1+ − + dx (7.4.13)
2
1 16 2 x

−−−−−−−−−−−
2 2
x 1 1
=∫ √ + + dx (7.4.14)
1
16 2 x2

−−−−−−−−−
2 2 2
x 1 x 1
=∫ √( + ) dx =∫ ( + ) dx (7.4.15)
1
4 x 1
4 x

2 2
x ∣
=( + ln x) ∣ (7.4.16)
8 ∣1

3
= + ln 2 ≈ 1.07 units. (7.4.17)
8

A graph of f is given in Figure 7.4.4; the portion of the curve measured in this problem is in bold.

The previous examples found the arc length exactly through careful choice of the functions. In general, exact answers are much
more difficult to come by and numerical approximations are necessary.

Example 7.4.3: Approximating arc length numerically

Find the length of the sine curve from x = 0 to x = π .


Solution
This is somewhat of a mathematical curiosity; in Example 5.4.3 we found the area under one "hump" of the sine curve is 2
square units; now we are measuring its arc length.
The setup is straightforward: f (x) = sin x and f ′
(x) = cos x . Thus
π
− −−− −−− −
L =∫ √ 1 + cos2 x dx. (7.4.18)
0

This integral cannot be evaluated in terms of elementary functions so we will approximate it with Simpson's Method with
n = 4.

−−−−−−− −
Figure 7.4.5 : A table of values of y = √1 + cos 2
x to evaluate a definite integral in Example 7.4.3.
− −−− −−− −
x √ 1 + cos2 x


0 √2
−−−
π/4 √3/2
(7.4.19)
π/2 1
−−−
3π/4 √3/2


π √2

−−−−−−− −
Figure ??? gives √1 + cos2
x evaluated at 5 evenly spaced points in [0, π]. Simpson's Rule then states that

7.4.4 https://math.libretexts.org/@go/page/4195
π
− −−− −−− − π −0 −−− −−−
2 – –
∫ √ 1 + cos x dx ≈ (√2 + 4 √3/2 + 2(1) + 4 √3/2 + √2) (7.4.20)
0
4⋅3

= 3.82918. (7.4.21)

Using a computer with n = 100 the approximation is L ≈ 3.8202; our approximation with n = 4 is quite good.

Surface Area of Solids of Revolution


We have already seen how a curve y = f (x) on [a, b] can be revolved around an axis to form a solid. Instead of computing its
volume, we now consider its surface area.

Figure 7.4.6 : Establishing the formula for surface area.


We begin as we have in the previous sections: we partition the interval [a, b] with n subintervals, where the i subinterval is th

[x , x
i ]. On each subinterval, we can approximate the curve y = f (x) with a straight line that connects f (x ) and f (x
i+1 ) as i i+1

shown in Figure 7.4.5a. Revolving this line segment about the x-axis creates part of a cone (called a frustum of a cone) as shown
in Figure 7.4.5b. The surface area of a frustum of a cone is
2π ⋅ length ⋅ average of the two radii R and r. (7.4.22)

The length is given by L; we use the material just covered by arc length to state that
−−−−−−−−

L ≈ √ 1 + f (ci ) dxi (7.4.23)

for some c in the i


i
th
subinterval. The radii are just the function evaluated at the endpoints of the interval. That is,
R = f (xi+1 ) and r = f (xi ). (7.4.24)

Thus the surface area of this sample frustum of the cone is approximately
f (xi ) + f (xi+1 ) −−−−−−−−−
′ 2
2π √ 1 + f (ci ) dxi . (7.4.25)
2

Since f is a continuous function, the Intermediate Value Theorem states there is some di in [ xi , xi+1 ] such that
f ( xi )+f ( xi+1 )
f (di ) =
2
; we can use this to rewrite the above equation as
−−−−−−−−−
′ 2
2πf (di )√ 1 + f (ci ) dxi . (7.4.26)

Summing over all the subintervals we get the total surface area to be approximately
n
−−−−−−−−−
′ 2
Surface Area ≈ ∑ 2πf (di )√ 1 + f (ci ) dxi , (7.4.27)

i=1

which is a Riemann Sum. Taking the limit as the subinterval lengths go to zero gives us the exact surface area, given in the
following Key Idea.

7.4.5 https://math.libretexts.org/@go/page/4195
Key Idea 28: Surface Area of a Solid of Revolution
Let f be differentiable on an open interval containing [a, b] where f is also continuous on [a, b].

1. The surface area of the solid formed by revolving the graph of y = f (x), where f (x) ≥ 0 , about the x-axis is
$$\text{Surface Area} = 2\pi\int_a^b f(x)\sqrt{1+f'(x)^2}\ dx.\]
2. The surface area of the solid formed by revolving the graph of y = f (x) about the y -axis, where a, b ≥ 0 , is
$$\text{Surface Area} = 2\pi\int_a^b x\sqrt{1+f'(x)^2}\ dx.\]

When revolving y = f (x) about the y -axis, the radii of the resulting frustum are x and x ; their average value is simply the
i i+1

midpoint of the interval. In the limit, this midpoint is just x. This gives the second part of Key Idea 28.

Example 7.4.4: Finding surface area of a solid of revolution

Find the surface area of the solid formed by revolving y = sin x on [0, π] around the x-axis, as shown in Figure 7.4.6.

Figure 7.4.7 : Revolving y = sin x on [0, π] about the x-axis.


Solution
The setup is relatively straightforward. Using Key Idea ??? , we have the surface area SA is:
π
−−−−−−−−
2
SA = 2π ∫ sin x √ 1 + cos x dx (7.4.28)
0

−−−−−−−− π
1 −1 2 ∣
= −2π (sinh (cos x) + cos x √ 1 + cos x) (7.4.29)

2 0
– −1
= 2π (√2 + sinh 1) (7.4.30)

2
≈ 14.42 units . (7.4.31)

The integration step above is nontrivial, utilizing an integration method called Trigonometric Substitution.
It is interesting to see that the surface area of a solid, whose shape is defined by a trigonometric function, involves both a
square root and an inverse hyperbolic trigonometric function.

Example 7.4.5: Finding surface area of a solid of revolution

Find the surface area of the solid formed by revolving the curve y = x on [0, 1] about:
2

1. the x-axis
2. the y -axis.

7.4.6 https://math.libretexts.org/@go/page/4195
Figure 7.4.8 : The solids used in Example 7.4.6.
Solution
1. The integral is straightforward to setup:
1 −−−−−−−−
2 2
SA = 2π ∫ x √ 1 + (2x) dx. (7.4.32)
0

Like the integral in Example ??? , this requires Trigonometric Substitution.


1
π − −−−− − ∣
3 2 −1
= (2(8 x + x)√ 1 + 4x − sinh (2x)) (7.4.33)
32 ∣
0

π – −1
= (18 √5 − sinh 2) (7.4.34)
32
2
≈ 3.81 units . (7.4.35)

The solid formed by revolving y = x around the x-axis is graphed in Figure 7.4.7a.
2

2. Since we are revolving around the y -axis, the "radius" of the solid is not f (x) but rather x. Thus the integral to compute the
surface area is:
1 −−−−−−−−
2
SA = 2π ∫ x √ 1 + (2x) dx. (7.4.36)
0

This integral can be solved using substitution. Set u = 1 + 4x ; the new bounds are u = 1 to u = 5 . We then have
2

5
π

= ∫ √u du (7.4.37)
4 1

π 2
3/2 5
= u ∣
∣1 (7.4.38)
4 3
π –
= (5 √5 − 1) (7.4.39)
6
2
≈ 5.33 units . (7.4.40)

The solid formed by revolving y = x about the y -axis is graphed in Figure 7.4.7b.
2

Our final example is a famous mathematical "paradox."

Example 7.4.6: The surface area and volume of Gabriel's Horn

Consider the solid formed by revolving y = 1/x about the x-axis on [1, ∞). Find the volume and surface area of this solid.
(This shape, as graphed in Figure 7.4.9, is known as "Gabriel's Horn" since it looks like a very long horn that only a
supernatural person, such as an angel, could play.)

7.4.7 https://math.libretexts.org/@go/page/4195
Figure 7.4.9 : A graph of Gabriel's Horn.
Solution
To compute the volume it is natural to use the Disk Method. We have:

1
V =π∫ dx (7.4.41)
2
1 x
b
1
= lim π ∫ dx (7.4.42)
2
b→∞
1 x
b
−1 ∣
= lim π ( )∣ (7.4.43)
b→∞ x ∣
1

1
= lim π (1 − ) (7.4.44)
b→∞ b

3
= π units . (7.4.45)

Gabriel's Horn has a finite volume of π cubic units. Since we have already seen that regions with infinite length can have a
finite area, this is not too difficult to accept.
We now consider its surface area. The integral is straightforward to setup:
∞ −−−−−−−
1
4
SA = 2π ∫ √ 1 + 1/x dx. (7.4.46)
1
x

−−−− −−−
Integrating this expression is not trivial. We can, however, compare it to other improper integrals. Since 1 < √1 + 1/x
4
on
[1, ∞), we can state that

∞ ∞ −−−−−−−
1 1 4
2π ∫ dx < 2π ∫ √ 1 + 1/x dx. (7.4.47)
1
x 1
x

By Key Idea 21, the improper integral on the left diverges. Since the integral on the right is larger, we conclude it also diverges,
meaning Gabriel's Horn has infinite surface area.
Hence the "paradox": we can fill Gabriel's Horn with a finite amount of paint, but since it has infinite surface area, we can
never paint it.
Somehow this paradox is striking when we think about it in terms of volume and area. However, we have seen a similar
paradox before, as referenced above. We know that the area under the curve y = 1/x on [1, ∞) is finite, yet the shape has an
2

infinite perimeter. Strange things can occur when we deal with the infinite.

A standard equation from physics is "Work = force × distance", when the force applied is constant. In the next section we learn
how to compute work when the force applied is variable.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

7.4.8 https://math.libretexts.org/@go/page/4195
Integrated by Justin Marshall.

This page titled 7.4: Arc Length and Surface Area is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory
Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

7.4.9 https://math.libretexts.org/@go/page/4195
7.5: Work
Work is the scientific term used to describe the action of a force which moves an object. When a constant force F

is applied to
move an object a distance d , the amount of work performed is
⃗ ⃗
W =F ⋅d. (7.5.1)

The SI unit of force is the Newton, (kg⋅m/s ) and the SI unit of distance is a meter (m). The fundamental unit of work is one
2

Newton-meter, or a joule (J). That is, applying a force of one Newton for one meter performs one joule of work. In Imperial units
(as used in the United States), force is measured in pounds (lb) and distance is measured in feet (ft), hence work is measured in ft-
lb.

Mass vs. Weight


Mass and weight are closely related, yet different, concepts. The mass m of an object is a quantitative measure of that object's
resistance to acceleration. The weight w of an object is a measurement of the force applied to the object by the acceleration of
gravity g .
Since the two measurements are proportional, w = m ⋅ g , they are often used interchangeably in everyday conversation. When
computing work, one must be careful to note which is being referred to. When mass is given, it must be multiplied by the
acceleration of gravity to reference the related force.

When force is constant, the measurement of work is straightforward. For instance, lifting a 200 lb object 5 ft performs
200 ⋅ 5 = 1000 ft-lb of work.

What if the force applied is variable? For instance, imagine a climber pulling a 200 ft rope up a vertical face. The rope becomes
lighter as more is pulled in, requiring less force and hence the climber performs less work.
In general, let F (x) be a force function on an interval [a, b]. We want to measure the amount of work done applying the force F
from x = a to x = b . We can approximate the amount of work being done by partitioning [a, b] into subintervals
a =x <x <⋯ <x
1 2 n+1 =b and assuming that F is constant on each subinterval. Let c be a value in the i
i subinterval
th

[x , x
i ]. Then the work done on this interval is approximately
i+1

Wi ≈ F (ci ) ⋅ (xi+1 − xi ) = F (ci ) Δxi ,

a constant force × the distance over which it is applied. The total work is
n n

W = ∑ Wi ≈ ∑ F (ci ) Δxi .

i=1 i=1

This, of course, is a Riemann sum. Taking a limit as the subinterval lengths go to zero give an exact value of work which can be
evaluated through a definite integral.

Key Idea 29: Work


Let F (x) be a continuous function on [a, b] describing the amount of force being applied to an object in the direction of travel
from distance x = a to distance x = b . The total work W done on [a, b] is
b

W =∫ F (x) dx. (7.5.2)


a

Example 7.5.1: Computing work performed - applying variable force

A 60m climbing rope is hanging over the side of a tall cliff. How much work is performed in pulling the rope up to the top,
where the rope has a mass of 66g/m? How much work is performed pulling a 60 m climbing rope up a cliff face, where the
rope has a mass of 66 g/m?
Solution

7.5.1 https://math.libretexts.org/@go/page/4196
We need to create a force function F (x) on the interval [0, 60]. To do so, we must first decide what x is measuring: it is the
length of the rope still hanging or is it the amount of rope pulled in? As long as we are consistent, either approach is fine. We
adopt for this example the convention that x is the amount of rope pulled in. This seems to match intuition better; pulling up
the first 10 meters of rope involves x = 0 to x = 10 instead of x = 60 to x = 50.
As x is the amount of rope pulled in, the amount of rope still hanging is 60 − x . This length of rope has a mass of 66 g/m, or
0.066 kg/m. The the mass of the rope still hanging is 0.066(60 − x) kg; multiplying this mass by the acceleration of gravity,

9.8 m/s , gives our variable force function


2

F (x) = (9.8)(0.066)(60 − x) = 0.6468(60 − x).

Thus the total work performed in pulling up the rope is


60

W =∫ 0.6468(60 − x) dx = 1, 164.24 J.
0

By comparison, consider the work done in lifting the entire rope 60 meters. The rope weights 60 × 0.066 × 9.8 = 38.808 N,
so the work applying this force for 60 meters is 60 × 38.808 = 2, 328.48 J. This is exactly twice the work calculated before
(and we leave it to the reader to understand why.)

Example 7.5.2: Computing work performed - applying variable force

Consider again pulling a 60 m rope up a cliff face, where the rope has a mass of 66 g/m. At what point is exactly half the work
performed?
Solution
From Example 7.5.1 we know the total work performed is 1, 164.24 J. We want to find a height h such that the work in
pulling the rope from a height of x = 0 to a height of x = h is 582.12, half the total work. Thus we want to solve the equation
h

∫ 0.6468(60 − x) dx = 582.12
0

for h .
h

∫ 0.6468(60 − x) dx = 582.12
0

h
2 ∣
(38.808x − 0.3234 x ) = 582.12

0

2
38.808h − 0.3234h = 582.12

2
−0.3234 h + 38.808h − 582.12 = 0.

Apply the Quadratic Formula.

h = 17.57 and 102.43

As the rope is only 60 m long, the only sensible answer is h = 17.57. Thus about half the work is done pulling up the first
17.5 m the other half of the work is done pulling up the remaining 42.43 m.

Example 7.5.3: Computing work performed: applying variable force

A box of 100 lb of sand is being pulled up at a uniform rate a distance of 50 ft over 1 minute. The sand is leaking from the box
at a rate of 1 lb/s. The box itself weighs 5 lb and is pulled by a rope weighing .2 lb/ft.
a. How much work is done lifting just the rope?
b. How much work is done lifting just the box and sand?
c. What is the total amount of work performed?
Solution

7.5.2 https://math.libretexts.org/@go/page/4196
a. We start by forming the force function F (x) for the rope (where the subscript denotes we are considering the rope). As in
r

the previous example, let x denote the amount of rope, in feet, pulled in. (This is the same as saying x denotes the height of
the box.) The weight of the rope with x feet pulled in is

Fr (x) = 0.2(50 − x) = 10 − 0.2x.

(Note that we do not have to include the acceleration of gravity here, for the weight of the rope per foot is given, not its
mass per meter as before.) The work performed lifting the rope is
50

Wr = ∫ (10 − 0.2x) dx = 250 ft-lb.


0

b. The sand is leaving the box at a rate of 1 lb/s. As the vertical trip is to take one minute, we know that 60 lb will have left
when the box reaches its final height of 50 ft. Again letting x represent the height of the box, we have two points on the line
that describes the weight of the sand: when x = 0 , the sand weight is 100 lb, producing the point (0, 100); when x = 50,
100−40
the sand in the box weighs 40 lb, producing the point (50, 40). The slope of this line is = −1.2 , giving the equation
0−50

of the weight of the sand at height x as w(x) = −1.2x + 100. The box itself weighs a constant 5 lb, so the total force
function is F (x) = −1.2x + 105 . Integrating from x = 0 to x = 50 gives the work performed in lifting box and sand:
b

50

Wb = ∫ (−1.2x + 105) dx = 3750 ft-lb.


0

c. The total work is the sum of W and W : 250 + 3750 = 4000 ft-lb. We can also arrive at this via integration:
r b

50

W =∫ (Fr (x) + Fb (x)) dx


0

50

=∫ (10 − 0.2x − 1.2x + 105) dx


0

50

=∫ (−1.4x + 115) dx
0

= 4000 ft-lb.

Hooke's Law and Springs


Hooke's Law states that the force required to compress or stretch a spring x units from its natural length is proportional to x; that is,
this force is F (x) = kx for some constant k . For example, if a force of 1 N stretches a given spring 2 cm, then a force of 5 N will
stretch the spring 10 cm. Converting the distances to meters, we have that stretching this spring 0.02 m requires a force of
F (0.02) = k(0.02) = 1 N, hence k = = 50 N/m.
1

0.02

Example 7.5.4: Computing work performed: stretching a spring

A force of 20 lb stretches a spring from a natural length of 7 inches to a length of 12 inches. How much work was performed in
stretching the spring to this length?
Solution
In many ways, we are not at all concerned with the actual length of the spring, only with the amount of its change. Hence, we
do not care that 20 lb of force stretches the spring to a length of 12 inches, but rather that a force of 20 lb stretches the spring
by 5 in. This is illustrated in Figure 7.5.1; we only measure the change in the spring's length, not the overall length of the
spring.

Figure 7.5.1 : Illustrating the important aspects of stretching a spring in computing work in Example 7.5.4 .

7.5.3 https://math.libretexts.org/@go/page/4196
Converting the units of length to feet, we have
5 5
F ( ) = k = 20 lb.
12 12

Thus k = 48 lb/ft and F (x) = 48x.


We compute the total work performed by integrating F (x) from x = 0 to x = 5

12
:
5/12

W =∫ 48x dx
0

5/12
2∣
= 24x

0

25
= ≈ 4.1667 ft-lb.
6

Pumping Fluids
Another useful example of the application of integration to compute work comes in the pumping of fluids, often illustrated in the
context of emptying a storage tank by pumping the fluid out the top. This situation is different than our previous examples for the
forces involved are constant. After all, the force required to move one cubic foot of water (about 62.4 lb) is the same regardless of
its location in the tank. What is variable is the distance that cubic foot of water has to travel; water closer to the top travels less
distance than water at the bottom, producing less work.
Figure 7.5.2 : Weight and Mass densities.
Fluid lb/ft3 kg/m3

Concrete 150 2400

Fuel Oil 55.46 890.13

Gasoline 45.93 737.22

Iodine 307 4927

Methanol 49.3 791.3

Mercury 844 1354

Milk 63.6-65.4 1020-1050

Water 62.4 1000

We demonstrate how to compute the total work done in pumping a fluid out of the top of a tank in the next two examples.

Example 7.5.5: Computing work performed: pumping fluids

A cylindrical storage tank with a radius of 10 ft and a height of 30 ft is filled with water, which weighs approximately 62.4 lb/ft
3
. Compute the amount of work performed by pumping the water up to a point 5 feet above the top of the tank.
Solution
We will refer often to Figure 7.5.3 which illustrates the salient aspects of this problem.

Figure 7.5.3 : Illustrating a water tank in order to compute the work required to empty it in Example 7.5.5.

7.5.4 https://math.libretexts.org/@go/page/4196
We start as we often do: we partition an interval into subintervals. We orient our tank vertically since this makes intuitive sense
with the base of the tank at y = 0 . Hence the top of the water is at y = 30, meaning we are interested in subdividing the y -
interval [0, 30] into n subintervals as

0 = y1 < y2 < ⋯ < yn+1 = 30.

Consider the work W of pumping only the water residing in the i subinterval, illustrated in Figure 7.5.3. The force required
i
th

to move this water is equal to its weight which we calculate as volume × density. The volume of water in this subinterval is
V = 10 πΔy ; its density is 62.4 lb/ft . Thus the required force is 6240πΔy lb.
2 3
i i i

We approximate the distance the force is applied by using any y -value contained in the i subinterval; for simplicity, we
th

arbitrarily use y for now (it will not matter later on). The water will be pumped to a point 5 feet above the top of the tank, that
i

is, to the height of y = 35 ft. Thus the distance the water at height y travels is 35 − y ft.
i i

In all, the approximate work W performed in moving the water in the i


i
th
subinterval to a point 5 feet above the tank is

Wi ≈ 6240πΔyi (35 − yi ).

To approximate the total work performed in pumping out all the water from the tank, we sum all the work Wi performed in
pumping the water from each of the n subintervals of [0, 30]:
n n

W ≈ ∑ Wi = ∑ 6240πΔyi (35 − yi ).

i=1 i=1

This is a Riemann sum. Taking the limit as the subinterval length goes to 0 gives
30

W =∫ 6240π(35 − y) dy
0

30
1 2 ∣
= 6240π (35y − y )
2 ∣
0

= 11, 762, 123 ft-lb

7
≈ 1.176 × 10 ft-lb.

We can "streamline'' the above process a bit as we may now recognize what the important features of the problem are. Figure 7.5.4
shows the tank from Example 7.5.5 without the i subinterval identified.
th

Figure 7.5.4 : A simplified illustration for computing work.


Instead, we just draw one differential element. This helps establish the height a small amount of water must travel along with the
force required to move it (where the force is volume × density).
We demonstrate the concepts again in the next examples.

Example 7.5.6: Computing work performed - pumping fluids

A conical water tank has its top at ground level and its base 10 feet below ground. The radius of the cone at ground level is 2 ft.
It is filled with water weighing 62.4 lb/ft and is to be emptied by pumping the water to a spigot 3 feet above ground level.
3

Find the total amount of work performed in emptying the tank.

7.5.5 https://math.libretexts.org/@go/page/4196
Solution
The conical tank is sketched in Figure 7.5.5. We can orient the tank in a variety of ways; we could let y = 0 represent the base
of the tank and y = 10 represent the top of the tank, but we choose to keep the convention of the wording given in the problem
and let y = 0 represent ground level and hence y = −10 represents the bottom of the tank. The actual "height'' of the water
does not matter; rather, we are concerned with the distance the water travels.

Figure 7.5.5 : A graph of the conical water tank in Example 7.5.6 .


The figure also sketches a differential element, a cross-sectional circle. The radius of this circle is variable, depending on y .
When y = −10 , the circle has radius 0; when y = 0 , the circle has radius 2. These two points, (−10, 0) and (0, 2), allow us to
find the equation of the line that gives the radius of the cross-sectional circle, which is r(y) = y + 2 . Hence the volume of
1

water at this height is V (y) = π( y + 2) dy , where dy represents a very small height of the differential element. The force
1

5
2

required to move the water at height y is F (y) = 62.4 × V (y) .


The distance the water at height y travels is given by h(y) = 3 − y . Thus the total work done in pumping the water from the
tank is
0
1 2
W =∫ 62.4π( y + 2 ) (3 − y) dy
5
−10

0
1 3 17 2 8
= 62.4π ∫ (− y − y − y + 12) dy
25 25 5
−10

220
= 62.2π ⋅ ≈ 14, 376 ft-lb.
3

Example 7.5.7: Computing work performed - pumping fluids

A rectangular swimming pool is 20 ft wide and has a 3 ft "shallow end'' and a 6 ft "deep end.'' It is to have its water pumped
out to a point 2 ft above the current top of the water.
The cross-sectional dimensions of the water in the pool are given in Figure 7.5.6; note that the dimensions are for the water,
not the pool itself. Compute the amount of work performed in draining the pool.

Figure 7.5.6 : The cross-section of a swimming pool filled with water in Example 7.5.7 .
Solution
For the purposes of this problem we choose to set y = 0 to represent the bottom of the pool, meaning the top of the water is at
y = 6.

7.5.6 https://math.libretexts.org/@go/page/4196
Figure 7.5.7 : Orienting the pool and showing differential elements for Example 7.5.7 .
Figure 7.5.7 shows the pool oriented with this y -axis, along with 2 differential elements as the pool must be split into two
different regions.
The top region lies in the y -interval of [3, 6], where the length of the differential element is 25 ft as shown. As the pool is 20 ft
wide, this differential element represents a this slice of water with volume V (y) = 20 ⋅ 25 ⋅ dy . The water is to be pumped to a
height of y = 8 , so the height function is h(y) = 8 − y . The work done in pumping this top region of water is
6

Wt = 62.4 ∫ 500(8 − y) dy = 327, 600 ft-lb.


3

The bottom region lies in the y -interval of [0, 3]; we need to compute the length of the differential element in this interval.
One end of the differential element is at x = 0 and the other is along the line segment joining the points (10, 0) and (15, 3).
The equation of this line is y = (x − 10) ; as we will be integrating with respect to y , we rewrite this equation as
3

y + 10 . So the length of the differential element is a difference of x-values: x = 0 and x = y + 10 , giving a length of
5 5
x =
3 3

y + 10 .
5
x =
3

Again, as the pool is 20 ft wide, this differential element represents a thin slice of water with volume
V (y) = 20 ⋅ ( y + 10) ⋅ dy ; the height function is the same as before at h(y) = 8 − y . The work performed in emptying this
5

part of the pool is


3
5
Wb = 62.4 ∫ 20( y + 10)(8 − y) dy = 299, 520 ft-lb.
3
0

The total work in emptying the pool is

W = Wb + Wt = 327, 600 + 299, 520 = 627, 120 ft-lb.

Notice how the emptying of the bottom of the pool performs almost as much work as emptying the top. The top portion travels
a shorter distance but has more water. In the end, this extra water produces more work.

The next section introduces one final application of the definite integral, the calculation of fluid force on a plate.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 7.5: Work is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et al. via
source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

7.5.7 https://math.libretexts.org/@go/page/4196
7.6: Fluid Forces
In the unfortunate situation of a car driving into a body of water, the conventional wisdom is that the water pressure on the doors
will quickly be so great that they will be effectively unopenable. (Survival techniques suggest immediately opening the door,
rolling down or breaking the window, or waiting until the water fills up the interior at which point the pressure is equalized and the
door will open. See Mythbusters episode #72 to watch Adam Savage test these options.) How can this be true? How much force
does it take to open the door of a submerged car? In this section we will find the answer to this question by examining the forces
exerted by fluids.
We start with pressure, which is related to force by the following equations:
Force
Pressure = ⇔ Force = Pressure × Area. (7.6.1)
Area

In the context of fluids, we have the following definition.

Definition 26: Fluid pressure


Let w be the weight-density of a fluid. The pressure p exerted on an object at depth d in the fluid is p = w ⋅ d .

We use this definition to find the force exerted on a horizontal sheet by considering the sheet's area.

Example 7.6.1: Computing fluid force


1. A cylindrical storage tank has a radius of 2 ft and holds 10 ft of a fluid with a weight-density of 50 lb/ft (Figure 7.6.1.)
3

What is the force exerted on the base of the cylinder by the fluid?

Figure 7.6.1 : A cylindrical tank in Example 7.6.1 .


1. A rectangular tank whose base is a 5 ft square has a circular hatch at the bottom with a radius of 2 ft. The tank holds 10 ft of
a fluid with a weight-density of 50 lb/ft . (Figure 7.6.2). What is the force exerted on the hatch by the fluid?
3

Figure 7.6.2 : A rectangular tank in Example 7.6.1 .


Solution
3
1. Using Definition 26, we calculate that the pressure exerted on the cylinder's base is w ⋅ d = 50 lb/ft × 10 ft = 500 lb/ft .
2

The area of the base is π ⋅ 2 = 4π ft . So the force exerted by the fluid is


2 2

F = 500 × 4π = 6283 lb.

Note that we effectively just computed the weight of the fluid in the tank.

7.6.1 https://math.libretexts.org/@go/page/4197
2. The dimensions of the tank in this problem are irrelevant. All we are concerned with are the dimensions of the hatch and
the depth of the fluid. Since the dimensions of the hatch are the same as the base of the tank in the previous part of this
example, as is the depth, we see that the fluid force is the same. That is, F = 6283 lb.
A key concept to understand here is that we are effectively measuring the weight of a 10 ft column of water above the hatch.
The size of the tank holding the fluid does not matter.

The previous example demonstrates that computing the force exerted on a horizontally oriented plate is relatively easy to compute.
What about a vertically oriented plate? For instance, suppose we have a circular porthole located on the side of a submarine. How
do we compute the fluid force exerted on it?
Pascal's Principle states that the pressure exerted by a fluid at a depth is equal in all directions. Thus the pressure on any portion of
a plate that is 1 ft below the surface of water is the same no matter how the plate is oriented. (Thus a hollow cube submerged at a
great depth will not simply be "crushed'' from above, but the sides will also crumple in. The fluid will exert force on all sides of the
cube.)
So consider a vertically oriented plate as shown in Figure 7.6.3 submerged in a fluid with weight-density w. What is the total fluid
force exerted on this plate? We find this force by first approximating the force on small horizontal strips.

Figure 7.6.3 : A thin, vertically oriented plate submerged in a fluid with weight density w.
Let the top of the plate be at depth b and let the bottom be at depth a . (For now we assume that surface of the fluid is at depth 0, so
if the bottom of the plate is 3 ft under the surface, we have a = −3 . We will come back to this later.) We partition the interval [a, b]
into n subintervals

a = y1 < y2 < ⋯ < yn+1 = b, (7.6.2)

with the i th
subinterval having length Δy . The force F exerted on the plate in the i
i i
th
subinterval is F
i = Pressure × Area.

The pressure is depth ×w. We approximate the depth of this thin strip by choosing any value d in [y , y ]; the depth is i i i+1

approximately −d . (Our convention has d being a negative number, so −d is positive.) For convenience, we let d be an
i i i i

endpoint of the subinterval; we let d = y .


i i

The area of the thin strip is approximately length × width. The width is Δy . The length is a function of some y -value c in the i
i i
th

subinterval. We state the length is ℓ(c ) . Thus


i

Fi = Pressure × Area

= −yi ⋅ w × ℓ(ci ) ⋅ Δyi .

To approximate the total force, we add up the approximate forces on each of the n thin strips:
n n

F = ∑ Fi ≈ ∑ −w ⋅ yi ⋅ ℓ(ci ) ⋅ Δyi . (7.6.3)

i=1 i=1

This is, of course, another Riemann Sum. We can find the exact force by taking a limit as the subinterval lengths go to 0; we
evaluate this limit with a definite integral.

key idea 30: fluid force on a vertically oriented plate


Let a vertically oriented plate be submerged in a fluid with weight-density w where the top of the plate is at y =b and the
bottom is at y = a . Let ℓ(y) be the length of the plate at y .
1. If y = 0 corresponds to the surface of the fluid, then the force exerted on the plate by the fluid is

7.6.2 https://math.libretexts.org/@go/page/4197
b

F =∫ w ⋅ (−y) ⋅ ℓ(y)dy. (7.6.4)


a

2. In general, let d(y) represent the distance between the surface of the fluid and the plate at y . Then the force exerted on the
plate by the fluid is
b

F =∫ w ⋅ d(y) ⋅ ℓ(y) dy. (7.6.5)


a

Example 7.6.2: Finding fluid force

Consider a thin plate in the shape of an isosceles triangle as shown in Figure 7.6.4 submerged in water with a weight-density
of 62.4 lb/ft . If the bottom of the plate is 10 ft below the surface of the water, what is the total fluid force exerted on this
3

plate?

Figure 7.6.4 : A thin plate in the shape of an isosceles triangle in Example 7.6.2.
Solution
We approach this problem in two different ways to illustrate the different ways Key Idea 30 can be implemented. First we will
let y = 0 represent the surface of the water, then we will consider an alternate convention.

Figure 7.6.5 : Sketching the triangular plate in Example 7.6.2 with the convention that the water level is at y=0.
1. We let y = 0 represent the surface of the water; therefore the bottom of the plate is at y = −10 . We center the triangle on
the y -axis as shown in Figure 7.6.5. The depth of the plate at y is −y as indicated by the Key Idea. We now consider the
length of the plate at y .

We need to find equations of the left and right edges of the plate. The right hand side is a line that connects the points
(0, −10) and (2, −6): that line has equation x = 1/2(y + 10). (Find the equation in the familiar y = mx + b format and

solve for x.) Likewise, the left hand side is described by the line x = −1/2(y + 10) . The total length is the distance
between these two lines: ℓ(y) = 1/2(y + 10) − (−1/2(y + 10)) = y + 10.

The total fluid force is then


−6

F =∫ 62.4(−y)(y + 10)dy
−10

176
= 62.4 ⋅ ≈ 3660.8 lb.
3

2. Sometimes it seems easier to orient the thin plate nearer the origin. For instance, consider the convention that the bottom of
the triangular plate is at (0, 0), as shown in Figure 7.6.6. The equations of the left and right hand sides are easy to find.
They are y = 2x and y = −2x , respectively, which we rewrite as x = 1/2y and x = −1/2y. Thus the length function is
ℓ(y) = 1/2y − (−1/2y) = y .

7.6.3 https://math.libretexts.org/@go/page/4197
As the surface of the water is 10 ft above the base of the plate, we have that the surface of the water is at y = 10. Thus the
depth function is the distance between y = 10 and y ; d(y) = 10 − y . We compute the total fluid force as:
4

F =∫ 62.4(10 − y)(y) dy
0

≈ 3660.8 lb.

Figure 7.6.6 : Sketching the triangular plate in Example 7.6.2 with the convention that the base of the triangle is at (0, 0).
The correct answer is, of course, independent of the placement of the plate in the coordinate plane as long as we are consistent.

Example 7.6.3: Finding fluid force

Find the total fluid force on a car door submerged up to the bottom of its window in water, where the car door is a rectangle 40''
long and 27'' high (based on the dimensions of a 2005 Fiat Grande Punto.)
Solution
The car door, as a rectangle, is drawn in Figure 7.6.7. Its length is 10/3 ft and its height is 2.25 ft. We adopt the convention
that the top of the door is at the surface of the water, both of which are at y = 0 . Using the weight-density of water of 62.4 lb/ft
3
, we have the total force as
0

F =∫ 62.4(−y)10/3dy
−2.25

=∫ −208ydy
−2.25

0
2∣
= −104y
∣−2.25

= 526.5 lb.

Most adults would find it very difficult to apply over 500 lb of force to a car door while seated inside, making the door
effectively impossible to open. This is counter-intuitive as most assume that the door would be relatively easy to open. The
truth is that it is not, hence the survival tips mentioned at the beginning of this section.

Figure 7.6.7 : Sketching a submerged car door in Example 7.6.2.

Example 7.6.4: Finding fluid force

An underwater observation tower is being built with circular viewing portholes enabling visitors to see underwater life. Each
vertically oriented porthole is to have a 3 ft diameter whose center is to be located 50 ft underwater. Find the total fluid force
exerted on each porthole. Also, compute the fluid force on a horizontally oriented porthole that is under 50 ft of water.

7.6.4 https://math.libretexts.org/@go/page/4197
Figure 7.6.8 : Measuring the fluid force on an underwater porthole in Example 7.6.4.
Solution
We place the center of the porthole at the origin, meaning the surface of the water is at y = 50 and the depth function will be
d(y) = 50 − y ; see Figure 7.6.8.

−−−−−−−−
The equation of a circle with a radius of 1.5 is x + y = 2.25 ; solving for x we have x = ±√2.25 − y , where the positive
2 2 2

square root corresponds to the right side of the circle and the negative square root corresponds to the left side of the circle.
−−−− −−−−
Thus the length function at depth y is ℓ(y) = 2√2.25 − y . Integrating on [−1.5, 1.5]we have:
2

1.5 −−−−−−−−
2
F = 62.4 ∫ 2(50 − y)√ 2.25 − y dy
−1.5

1.5 −−−−−−−− −−−−−−−−


2 2
= 62.4 ∫ (100 √ 2.25 − y − 2y √ 2.25 − y )dy
−1.5

1.5 −−−−−−−− 1.5 −−−−−−−−


2 2
= 6240 ∫ (√ 2.25 − y )dy − 62.4 ∫ (2y √ 2.25 − y )dy.
−1.5 −1.5

The second integral above can be evaluated using Substitution. Let u = 2.25 − y with du = −2y dy . The new bounds are:
2

u(−1.5) = 0 and u(1.5) = 0; the new integral will integrate from u = 0 to u = 0 , hence the integral is 0.

The first integral above finds the area of half a circle of radius 1.5, thus the first integral evaluates to
6240 ⋅ π ⋅ 1.5 /2 = 22, 054. Thus the total fluid force on a vertically oriented porthole is 22, 054 lb.
2

Finding the force on a horizontally oriented porthole is more straightforward:


2
F = Pressure × Area = 62.4 ⋅ 50 × π ⋅ 1.5 = 22, 054 lb. (7.6.6)

That these two forces are equal is not coincidental; it turns out that the fluid force applied to a vertically oriented circle whose
center is at depth d is the same as force applied to a horizontally oriented circle at depth d .

We end this chapter with a reminder of the true skills meant to be developed here. We are not truly concerned with an ability to find
fluid forces or the volumes of solids of revolution. Work done by a variable force is important, though measuring the work done in
pulling a rope up a cliff is probably not.
What we are actually concerned with is the ability to solve certain problems by first approximating the solution, then refining the
approximation, then recognizing if/when this refining process results in a definite integral through a limit. Knowing the formulas
found inside the special boxes within this chapter is beneficial as it helps solve problems found in the exercises, and other
mathematical skills are strengthened by properly applying these formulas. However, more importantly, understand how each of
these formulas was constructed. Each is the result of a summation of approximations; each summation was a Riemann sum,
allowing us to take a limit and find the exact answer through a definite integral.
The next chapter addresses an entirely different topic: sequences and series. In short, a sequence is a list of numbers, where a series
is the summation of a list of numbers. These seemingly-simple ideas lead to very powerful mathematics.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

7.6.5 https://math.libretexts.org/@go/page/4197
This page titled 7.6: Fluid Forces is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et al.
via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

7.6.6 https://math.libretexts.org/@go/page/4197
7.E: Applications of Integration (Exercises)
7.1: Area Between Curves
Terms and Concepts
1. T/F: The are between curves is always positive.
2. T/F: Calculus can be used to find the area of basic geometric shapes.
3. In your own words, describe how to find the total area enclosed by y = f (x) and y = g(x) .

Problems
In Exercises 4-10, find the area of the shaded region in the given graph.
4.

5.

6.

7.

7.E.1 https://math.libretexts.org/@go/page/4770
8.

9.

10.

In Exercises 11-16, find the total area enclosed by the functions f and g .
11. f (x) = 2x 2
+ 5x − 3, g(x) = x
2
+ 4x − 1

12. f (x) = x2
− 3x + 2, g(x) = −3x + 3

13. f (x) = sin x, g(x) = 2x/π

14.. f (x) = x 3
− 4x
2
+ x − 1, g(x) = −x
2
+ 2x − 4


15. f (x) = x, g(x) = √x

16. f (x) = −x 3
+ 5x
2
+ 2x + 1, g(x) = 3 x
2
+x +3

17. The functions f (x) = cos(2x) and g(x) = sin x intersect infinitely many times, forming an infinite number of repeated,
enclosed regions. Find the areas of these regions.
In Exercises 18-22, find the area of the enclosed region in two ways:
1. by treating the boundaries as functions of x, and
2. by treating the boundaries as functions of y.

7.E.2 https://math.libretexts.org/@go/page/4770
18.

19.

20.

21.

22.

In Exercises 23-26, find the are triangle formed by the given three points.
23. (1, 1), (2, 3) and (3, 3)

24. (−1, 1), (1, 3) and (2, −1)

25. (1, 1), (3, 3) and (3, 3)

7.E.3 https://math.libretexts.org/@go/page/4770
26. (0, 0), (2, 5) and (5, 2)

27. Use the Trapezoidal Rule to approximate the area of the pictured lake whose lengths, in hundreds of feet, are measured in 100-
foot increments.

28. Use Simpson's Rule to approximate the area of the pictured lake whose lengths, in hundreds of feet, are measured in 200-foot
increments.

7.2: Volume by Cross-Sectional Area: Disk and Washer Methods


Terms and Concepts
1. T/F: A solid of revolution is formed by revolving a shape around an axis.
2. In your own words, explain how the Disk and Washer Methods are related.
3 Explain how the units of volume are found in the integral of Theorem 54: if A(x) has units of in , how does
2
∫ A(x) dx have
units of in ?
3

Problems
In Exercises 4-7, a region of the Cartesian plane is shaded. Use the Disk/Washer Method to find the volume of the solid of
revolution formed by revolving the region about the x-axis.
4.

7.E.4 https://math.libretexts.org/@go/page/4770
5.

6.

7.

In Exercises 8-11, a region of the Cartesian plane is shaded. Use the Disk/Washer Method to find the volume of the solid of
revolution formed by revolving the region about the y-axis.
8.

9.

7.E.5 https://math.libretexts.org/@go/page/4770
10.

(Hint: Integration By Parts will be necessary, twice. First let u = arccos


2
x , then let u = arccos x.)
11.

In Exercises 12-17, a region of the Cartesian plane is described. Use the Disk/Washer Method to find the volume of the solid
of revolution formed by rotating the region about each of the given axes.
12. Region bounded by: y = √−
x, y = 0 and x = 1.

Rotate about:
(a) the x-axis
(b) y = 1
(c) the y-axis
(d) x = 1
13. Region bounded by: y = 4 − x 2
and y = 0.

Rotate about:
(a) the x-axis
(b) y = 4
(c) y = −1
(d) x = 2
14. The triangle with vertices (1, 1), (1, 2) and (2, 1).

Rotate about:
(a) the x-axis
(b) y = 2
(c) the y-axis
(d) x = 1
15. Region bounded by: y = y = x 2
− 2x + 2, and y = 2x − 1.

Rotate about:
(a) the x-axis
(b) y = 1
(c) y = 5
−−−−−
16. Region bounded by: y = 1/√x 2
+ 1 , x = −1, x = 1 and the x-axis.

Rotate about:
(a) the x-axis
(b) y = 1
(c) y = −1

7.E.6 https://math.libretexts.org/@go/page/4770
17. Region bounded by: y = 2x, y = x and x = 2.

Rotate about:
(a) the x-axis
(b) y = 4
(c) the y-axis
(d) x = 2
In Exercises 18-21, a solid is described. Orient the solid along the x-axis such that a cross-sectional area function A(x) can
be obtained, then apply Theorem 54 to find the volume of the solid.
18. A right circular cone with height of 10 and base radius of 5.

19. A skew right circular cone with height of 10 and base radius of 5. (Hint: all cross-sections are circles.)

20. A right triangle cone with height of 10 and whose base is a right, isosceles triangle with side length 4.

21. A solid with length 10 with a rectangular base and triangular top, wherein one end is a square with side length 5 and the other
end is a triangle with base and height of 5.

7.3: The Shell Method


Terms and Concepts
1. T/F: A solid of revolution is formed by revolving a shape around an axis.
2. T/F: The Shell Method can only be used when the Washer Method fails.
3. T/F: The Shell Method works by integrating cross-sectional areas of a solid.
4. T/F: When finding the volume of a solid of revolution that was revolved around a vertical axis, the Shell Method integrates with
respect to x.

Problems
In Exercises 5-8, a region of the Cartesian plane is shaded. Use the Shell Method to find the volume of the solid of revolution
formed by revolving the region about the y-axis.

7.E.7 https://math.libretexts.org/@go/page/4770
5.

6.

7.

8.

In Exercises 9-12, a region of the Cartesian plane is shaded. Use the Shell Method to find the volume of the solid of
revolution formed by revolving the region about the x-axis.
9.

7.E.8 https://math.libretexts.org/@go/page/4770
10.

11.

12.

In Exercises 13-18, a region of the Cartesian plane is described. Use the Shell Method to find the volume of the solid of revolution
formed by rotating the region about each of the given axes.
13. Region bounded by: y = √−
x, y = 0 and x = 1.

Rotate about:
(a) the x-axis
(b) y = 1
(c) the y-axis
(d) x = 1
14. Region bounded by: y = 4 − x 2
and y = 0.

Rotate about:
(a) x = 2
(b) x = −2
(c) the x-axis
(d) y = 4
15. The triangle with vertices (1, 1), (1, 2) and (2, 1).

Rotate about:
(a) the x-axis
(b) x = 1
(c) the x-axis
(d) y = 2
16. Region bounded by: y = y = x 2
− 2x + 2, and y = 2x − 1.

Rotate about:
(a) the x-axis

7.E.9 https://math.libretexts.org/@go/page/4770
(b) x = 1
(c) x = −1
−−−−−
17. Region bounded by: y = 1/√x 2
+ 1 , x = 1 and the x and y-axis.

Rotate about:
(a) the x-axis
(b) x = 1
18. Region bounded by: y = 2x, y = x and x = 2.

Rotate about:
(a) the y-axis
(b) x = 2
(c) the x-axis
(d) y = 4

7.4: Arc Length and Surface Area


Terms and Concepts
1. T/F: The integral formula for computing Arc Length was found by first approximating arc length with straight line segments.
2. T/F: The integral formula for computing Arc Length includes a square-root, meaning the integration is probably easy.

Problems
In Exercises 3-12, find the arc length of the function on the given interval.
3. f (x) = x on [0, 1].

4. f (x) = √8x on [−1, 1].
5. f (x) = 1

3
x
3/2
−x
1/2
on [0, 1].

6. f (x) = 1

12
x
3
+
x
1
on [1, 4].


7. f (x) = 2x 3/2

1

6
√x on [0, 9].

8. f (x) = cosh x on [− ln 2, ln 2].


9. f (x) = 1

2
(e
2
+e
−x
) on [0, ln 5].

10. f (x) = 1

12
x
5
+
1

5x3
on [0.1, 1].

11. f (x) = ln(sin x) on [π/6, π/2].


12. f (x) = ln(cos x) on [0, π/4].
In Exercises 13-20, set up the integral to compute the arc length of the function on the given interval. Do no evaluate the
integral.
13. f (x) = x 2
on [0, 1].

14. f (x) = x 10
on [0, 1].


15. f (x) = √x on [0, 1].

16. f (x) = ln x on [1, e].


−−−−−
17. f (x) = √1 − x 2
on [−1, 1]. (Note: this describes the top half of a circle with radius 1.)
−−−− − −−
18. 2
f (x) = √1 − x /9 on [−3, 3] . (Note: this describes the top half of an eclipse with a major axis of length 6 and a minor axis
of length 2.)
19. f (x) = 1

x
on [1, 2] .
20. f (x) = sec x on [−π/4, π/4].

7.E.10 https://math.libretexts.org/@go/page/4770
In Exercises 21-28, use Simpson's Rule, with n = 4 , to approximate the arc length of the function on the given interval.
Note: these are the same problems as in Exercises 13-20.
21. f (x) = x 2
on [0, 1].

22. f (x) = x 10
on [0, 1].


23. f (x) = √x on [0, 1]. (Note: f ′
(x) is not defined at x = 0 .)
24. f (x) = ln x on [1, e].
−−−−−
25. f (x) = √1 − x 2
on [−1, 1]. (Note: f ′
(x) is not defined at the endpoints.)
−−−− − −−
26. 2
f (x) = √1 − x /9 on [−3, 3] . (Note: f ′
(x) is not defined at the endpoints.)
27. f (x) = 1

x
on [1, 2] .
28. f (x) = sec x on [−π/4, π/4].
In Exercises 29-33, find the surface area of the described solid of revolution.
29. The solid formed by revolving y = 2x on [0, 1] about the x-axis.
30. The solid formed by revolving y = x 2
on [0, 1] about the y-axis.
31. The solid formed by revolving y = x 3
on [0, 1] about the x-axis.

32. The solid formed by revolving y = √x on [0, 1] about the x-axis.
−−−− −
33. The solid formed by revolving y = √1 − x
2
on [−1, 1] about the x-axis.

7.5: Work
Terms and Concepts
1. What are the typical units of work?
2. If a man has a mass of 80kg on Earth, will his mass on the moon be bigger, smaller, or the same?
3. If a woman weights 130 lb on Earth, will her weight on the moon be bigger, smaller, or the same?

Problems
4. A 100 ft rope, weighing 0.1 lb/ft, hangs over the edge of a tall building.
(a) How much work is done pulling the entire rope to the top of the building?
(b) How much rope is pulled in when half of the total work is done?
5. A 50 m rope, with a mass density of 0.2 kg/m, hangs over the edge of a tall building.
(a) How much work is done pulling the entire rope to the top of the building?
(b) How much work is done pulling in the first 20 m?
6. A rope of length l ft hangs over the edge of tall cliff. (Assume the cliff is taller than the length of the rope.) The rope has a
weight density of d lb/ft.
(a) How much work is done pulling the entire rope to the top of the cliff?
(b) What percentage of the total work is done pulling in the first half of the rope?
(c) How much rope is pulled in when half of the total work is done?
7. A 20 m rope with mass density of 0.5 kg/m hangs over the edge of a 10 m building. How much work is done pulling the rope to
the top?
8. A crane lifts a 2000 lb load vertically 30 ft with a 1" cable weighing 1.68 lb/ft.
(a) How much work is done lifting the cable alone?
(b) How much work is done lifting the load alone?
(c) Could one conclude that the work done lifting the cable is negligible compared to the work done lifting the load?
9. A 100 lb bag of sand is lifted uniformly 120 ft in one minute. Sand leaks from the bag at a rate of 1/4 lb/s. What is the total work
done in lifting the bag?

7.E.11 https://math.libretexts.org/@go/page/4770
10. A box weighing 2 lb lifts 10 lb of sand vertically 50 ft. A crack in the box allows the sand to leak out such that 9 lb of sand is in
the box at the end of the trip. Assume the sand leaked out at a uniform rate. What is the total work done in lifting the box and sand?
11. A force of 1000 lb compresses a spring 3 in. How much work is performed in compressing the spring?
12. A force of 2 N stretches a spring 5 cm. How much work is performed in stretching the spring?
13. A force of 50 lb compresses a spring from a natural length of 18 in to 12 in. How much work is performed in compressing the
spring?
14. A force of 20 lb stretches a spring from a natural length of 6 in to 8 in. How much work is performed in stretching the spring?
15. A force of 7 N stretches a spring from a natural length of 11 cm to 21 cm. How much work is performed in stretching the spring
from a length of 16 cm to 21 cm?
16. A force of f N stretches a spring d m from its natural length. How much work is performed in stretching the spring?
17. A 20 lb weight is attached to a spring. The weight rests on the spring, compressing the spring from a natural length of 1 ft to 6
in.
How much work is done in lifting the box 1.5 ft (i.e, the spring will be stretched 1 ft beyond its natural length)?
18. A 20 lb weight is attached to a spring. The weight rests on the spring, compressing the spring from a natural length of 1 ft to 6
in.
How much work is done in lifting the box 6 in (i.e, bringing the spring back to its natural length)?
19. A 5 m tall cylindrical tank with radius of 2 m is filled with 3 m of gasoline, with a mass density of 737.22 kg/m . Compute the
3

total work performed in pumping all the gasoline to the top of the tank.
20. A 6 ft cylindrical tank with a radius of 3 ft is filled with water, which has a weight density of 62.4 lb/ft . The water is to be
3

pumped to a point 2 ft above the top of the tank.


(a) How much work is performed in pumping all the water from the tank?
(b) How much work is performed in pumping 3 ft of water from the tank?
(c) At what point is 1/2 of the total work done?
21. A gasoline tanker is filled with gasoline with a weight density of 45.93 lb/ft . The dispensing value at the base is jammed shut,
3

forcing the operator to empty the tank via pumping the gas to a point 1 ft above the top of the tank. Assume the tank is a perfect
cylinder, 20 ft long with a diameter of 7.5 ft. How much work is performed in pumping all the gasoline from the tank?
22. A fuel oil storage tank is 10 ft deep with trapezoidal sides, 5 ft at the top of the 2 ft at the bottom, and is 15 ft wide (see diagram
below). Given that fuel oil weighs 55.46 lb/ft , find the work performed in pumping all the oil from the tank to a point 3 ft above
3

the top of the tank.

23. A conical tank is 5 m deep with a top radius of 3 m. (This is similar to Example 224.) The tank is filled with pure water, with a
mass density of 1000 kg/m . 3

(a) Find the work performed in pumping all the water to the top of the tank.
(b) Find the work performed in pumping the top 2.5 m of water to the top of the tank.
(c) Find the work performed in pumping the top half of the water, by volume, to the top of the tank.
24. A water tank has the shape of a truncated cone, with dimensions given below, and is filled with water with a weight density of
62.4 lb/ft . Find the work performed in pumping all water to a point 1 ft above the top of the tank.
3

7.E.12 https://math.libretexts.org/@go/page/4770
25. A water tank has the shape of an inverted pyramid, with dimensions given below, and is filled with water with a mass density of
1000 kg/m . Find the work performed in pumping all water to a point 5 m above the top of the tank.
3

26. A water tank has the shape of a truncated, inverted pyramid, with dimensions given blow, and is filled with water with a mass
density of 1000 kg/m . Find the work performed in pumping all water to a point 1 m above the top of the tank.
3

7.6: Fluid Forces


Terms and Concepts
1. State in your own words Pascal's Principle.
2. State in your own words how pressure is different from force.

Problems
In Exercises 3-12, find the fluid force exerted on the given plate, submerged in water with a weight density of 62.4 lb/ft . 3

3.

4.

7.E.13 https://math.libretexts.org/@go/page/4770
5.

6.

7.

8.

9.

7.E.14 https://math.libretexts.org/@go/page/4770
10.

11.

12.

In Exercises 13-18, the side of a container is pictured. Find the fluid force exerted on this plate when the container is full of:
1. water, with a weight density of 62.4 lb/ft
3

2. concrete, with a weight density of 150 lb/ft .


3

13.

14.

7.E.15 https://math.libretexts.org/@go/page/4770
15.

16.

17.

18.

19. How deep must the center of a vertically oriented circular plate with a radius of 1 ft be submerged in water, with a weight
density of 62.4 lb/ft , for the fluid force on the plate to reach 1,000 lb?
3

20. How deep must the center of a vertically oriented square plate with a side length of 2 ft be submerged in water, with a weight
density of 62.4 lb/ft , for the fluid force on the plate to reach 1,000 lb?
3

This page titled 7.E: Applications of Integration (Exercises) is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated
by Gregory Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is
available upon request.

7.E.16 https://math.libretexts.org/@go/page/4770
CHAPTER OVERVIEW
8: Sequences and Series
This chapter introduces sequences and series, important mathematical constructions that are useful when solving a large variety of
mathematical problems. The content of this chapter is considerably different from the content of the chapters before it. While the
material we learn here definitely falls under the scope of "calculus,'' we will make very little use of derivatives or integrals. Limits
are extremely important, though, especially limits that involve infinity.
One of the problems addressed by this chapter is this: suppose we know information about a function and its derivatives at a point,
such as f (1) = 3 , f (1) = 1 , f (1) = −2 , f (1) = 7 , and so on. What can I say about f (x) itself? Is there any reasonable
′ ′′ ′′′

approximation of the value of f (2)? The topic of Taylor Series addresses this problem, and allows us to make excellent
approximations of functions when limited knowledge of the function is available.
8.1: Sequences
8.2: Infinite Series
8.3: Integral and Comparison Tests
8.4: Ratio and Root Tests
8.5: Alternating Series and Absolute Convergence
8.6: Power Series
8.7: Taylor Polynomials
8.8: Taylor Series
8.E: Applications of Sequences and Series (Exercises)

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 8: Sequences and Series is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman
et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

1
8.1: Sequences
We commonly refer to a set of events that occur one after the other as a sequence of events. In mathematics, we use the word
sequence to refer to an ordered set of numbers, i.e., a set of numbers that "occur one after the other.''
For instance, the numbers 2, 4, 6, 8, ..., form a sequence. The order is important; the first number is 2, the second is 4, etc. It seems
natural to seek a formula that describes a given sequence, and often this can be done. For instance, the sequence above could be
described by the function a(n) = 2n , for the values of n = 1, 2, … To find the 10 term in the sequence, we would compute th

a(10). This leads us to the following, formal definition of a sequence.

Definition 27: sequences, range and terms


A sequence is a function a(n) whose domain is N.
The range of a sequence is the set of all distinct values of a(n) .
The terms of a sequence are the values a(1), a(2), ..., which are usually denoted with subscripts as a , a , .... 1 2

A sequence a(n) is often denoted as {a .


n}

Notation: We use N to describe the set of natural numbers, that is, the integers 1, 2, 3, ...

Definition: factorial
The expression 3! refers to the number 3 ⋅ 2 ⋅ 1 = 6 . In general,
n! = n ⋅ (n − 1) ⋅ (n − 2) ⋯ 2 ⋅ 1 (8.1.1)

where n is a natural number. We define 0! = 1 . While this does not immediately make sense, it makes many mathematical
formulas work properly.

Example 8.1.1: Listing terms of a sequence

List the first four terms of the following sequences.


n

1. {a n} ={
3

n!
}

2. {a n} = {4 + (−1 ) }
n

n( n+1) /2
(−1)
3. {a n} ={
n
2
}

Solution
1 2 3 4

1. a = = 3;
1
3

1!
a = = 2 ;
3

2!
a =
9

2
= ; 3a =
3
=
3!
9

2
4
3

4!
27

We can plot the terms of a sequence with a scatter plot. The "x''-axis is used for the values of n , and the values of the terms
are plotted on the y -axis. To visualize this sequence, see Figure 8.1(a).
2. a = 4 + (−1) = 3;
1
1
a = 4 + (−1 ) = 5;
2
2
a = 4 + (−1 ) = 3; a = 4 + (−1 ) = 5
3
3
. Note that the range
4
4

of this sequence is finite, consisting of only the values 3 and 5. This sequence is plotted in Figure 8.1(b).
3. (−1)
1(2)/2
(−1) 1
2(3)/2

a1 = = −1; a2 = =− (8.1.2)
2 2
1 2 4
3(4)/2 4(5)/2
(−1) 1 (−1) 1
a3 = = a4 = = ; (8.1.3)
2 2
3 9 4 16
5(6)/2
(−1) 1
a5 = =− (8.1.4)
2
5 25

.
We gave one extra term to begin to show the pattern of signs is "−, −, +, +, −, −, …, due to the fact that the exponent of
−1 is a special quadratic. This sequence is plotted in Figure 8.1(c).

8.1.1 https://math.libretexts.org/@go/page/4199
Figure 8.1: Plotting sequences in Example 8.1.1.

Example 8.1.2: Determining a formula for a sequence

Find the n th
term of the following sequences, i.e., find a function that describes each of the given sequences.
1. 2, 5, 8, 11, 14, …
2. 2, −5, 10, −17, 26, −37, …
3. 1, 1, 2, 6, 24, 120, 720, …
4. , , , , , …
5

2
5

2
15

8
5

4
25

32

Solution
We should first note that there is never exactly one function that describes a finite set of numbers as a sequence. There are
many sequences that start with 2, then 5, as our first example does. We are looking for a simple formula that describes the
terms given, knowing there is possibly more than one answer.
1. Note how each term is 3 more than the previous one. This implies a linear function would be appropriate:
a(n) = a = 3n + b for some appropriate value of b . As we want a = 2 , we set b = −1 . Thus a = 3n − 1 .
n 1 n

2. First notice how the sign changes from term to term. This is most commonly accomplished by multiplying the terms by
either (−1) or (−1) . Using (−1) multiplies the odd terms by (−1); using (−1)
n n+1 n
multiplies the even terms by
n+1

(−1). As this sequence has negative even terms, we will multiply by (−1) . n+1

After this, we might feel a bit stuck as to how to proceed. At this point, we are just looking for a pattern of some sort: what
do the numbers 2, 5, 10, 17, etc., have in common? There are many correct answers, but the one that we'll use here is that
each is one more than a perfect square. That is, 2 = 1 + 1 , 5 = 2 + 1 , 10 = 3 + 1 , etc. Thus our formula is
1 2 2

(n + 1) .
n+1 2
a = (−1 )
n

3. One who is familiar with the factorial function will readily recognize these numbers. They are 0!, 1!, 2!, 3!, etc. Since our
sequences start with n = 1 , we cannot write a = n! , for this misses the 0! term. Instead, we shift by 1, and write
n

a = (n − 1)! .
n

8.1.2 https://math.libretexts.org/@go/page/4199
4. This one may appear difficult, especially as the first two terms are the same, but a little ``sleuthing'' will help. Notice how
the terms in the numerator are always multiples of 5, and the terms in the denominator are always powers of 2. Does
something as simple as a = work?
n
5n
n
2

When n = 1 , we see that we indeed get 5/2 as desired. When n = 2 , we get 10/4 = 5/2. Further checking shows that this
formula indeed matches the other terms of the sequence.

A common mathematical endeavor is to create a new mathematical object (for instance, a sequence) and then apply previously
known mathematics to the new object. We do so here. The fundamental concept of calculus is the limit, so we will investigate what
it means to find the limit of a sequence.

Definition 28 LIMIT OF A SEQUENCE, CONVERGENT, DIVERGENT


Let { an } be a sequence and let L be a real number. Given any ϵ > 0 , if an m can be found such that | an − L| < ϵ for all
n >m , then we say the limit of {a }, as n approaches infinity, is L, denoted
n

lim an = L. (8.1.5)
n→∞

If lim an exists, we say the sequence converges; otherwise, the sequence diverges.
n→∞

This definition states, informally, that if the limit of a sequence is L, then if you go far enough out along the sequence, all
subsequent terms will be really close to L. Of course, the terms "far enough'' and "really close'' are subjective terms, but hopefully
the intent is clear.
This definition is reminiscent of the ϵ--δ proofs of Chapter 1. In that chapter we developed other tools to evaluate limits apart from
the formal definition; we do so here as well.

tHEOREM 55: LIMIT OF A SEQUENCE

Let {a } be a sequence and let f (x) be a function whose domain contains the positive real numbers where f (n) = a for all
n n

n in N .

Theorem 55 allows us, in certain cases, to apply the tools developed in Chapter 1 to limits of sequences. Note two things not stated
by the theorem:
1. If lim f (x) does not exist, we cannot conclude that lim an does not exist. It may, or may not, exist. For instance, we can
x→∞ n→∞

define a sequence {a } = {cos(2πn)}. Let f (x) = cos(2πx). Since the cosine function oscillates over the real numbers, the
n

limit lim f (x) does not exist.


x→∞

However, for every positive integer n , cos(2πn) = 1, so lim an = 1 .


n→∞

2. If we cannot find a function f (x) whose domain contains the positive real numbers where f (n) = a for all n in N, we cannot
n

conclude lim a does not exist. It may, or may not, exist.


n
n→∞

Example 8.1.3: Determining convergence/divergence of a sequence

Determine the convergence or divergence of the following sequences.


2

1. {a n} ={
3 n −2n+1

n2 −1000
}

2. {a n} = {cos n}
n
(−1)
3. {a n} ={
n
}

Solution

8.1.3 https://math.libretexts.org/@go/page/4199
2

1. Using Theorem 11, we can state that lim


3 x −2x+1
2
x −1000
=3 . (We could have also directly applied l'H\^opital's Rule.) Thus the
x→∞

sequence {a } converges, and its limit is 3. A scatter plot of every 5 values of a is given in Figure 8.2 (a). The values of
n n

a
n vary widely near n = 30 , ranging from about −73 to 125, but as n grows, the values approach 3.
2. The limit lim cos x does not exist, as cos x oscillates (and takes on every value in [−1, 1] infinitely many times). Thus we
x→∞

cannot apply Theorem 55.

The fact that the cosine function oscillates strongly hints that cos n, when n is restricted to N, will also oscillate. Figure 8.2
(b), where the sequence is plotted, shows that this is true. Because only discrete values of cosine are plotted, it does not
bear strong resemblance to the familiar cosine wave.

We conclude that lim an does not exist.


n→∞

3. We cannot actually apply Theorem 55 here, as the function f (x) = (−1) /x is not well defined. (What does (−1)
x √2

mean? In actuality, there is an answer, but it involves complex analysis, beyond the scope of this text.) So for now we say
that we cannot determine the limit. (But we will be able to very soon.) By looking at the plot in Figure 8.2 (c), we would
like to conclude that the sequence converges to 0. That is true, but at this point we are unable to decisively say so.

Figure 8.2: Scatter plots of the sequences in Example 8.1.2.

It seems that {(−1) n


/n} converges to 0 but we lack the formal tool to prove it. The following theorem gives us that tool.

THEOREM 56: ABSOLUTE VALUE THEOREM

Let {a n} be a sequence. If lim | an | = 0 , then lim an = 0


n→∞ n→∞

8.1.4 https://math.libretexts.org/@go/page/4199
Example 8.1.4: Determining the convergence/divergence of a sequence

Determine the convergence or divergence of the following sequences.


n
(−1)
1. {an} ={
n
}
n
(−1 ) (n+1)
2. {an} ={
n
}

Solution
1. This appeared in Example 8.1.1. We want to apply Theorem 56, so consider the limit of {|a :
n |}

n
∣ (−1) ∣
lim | an | = lim ∣ ∣
n→∞ n→∞ ∣ n ∣

1
= lim
n→∞ n

= 0.

Since this limit is 0, we can apply Theorem 56 and state that lim an = 0 .
n→∞

2. Because of the alternating nature of this sequence (i.e., every other term is multiplied by −1), we cannot simply look at the
x
(−1 ) (x+1)
limit lim
x
. We can try to apply the techniques of Theorem 56:
x→∞

n
∣ (−1 ) (n + 1) ∣
lim | an | = lim ∣ ∣
n→∞ n→∞ ∣ n ∣

n+1
= lim
n→∞ n

= 1.

We have concluded that when we ignore the alternating sign, the sequence approaches 1. This means we cannot apply
Theorem 56; it states the the limit must be 0 in order to conclude anything.

Since we know that the signs of the terms alternate and we know that the limit of |a | is 1, we know that as n approaches
n

infinity, the terms will alternate between values close to 1 and −1, meaning the sequence diverges. A plot of this sequence
is given in Figure 8.3.

Figure 8.3: A plot of a sequence in Example 8.1.2, part 2.

We continue our study of the limits of sequences by considering some of the properties of these limits.

THEOREM 57: Properties of the Limits of Sequences


Let {a n} and {b n} be sequences such that lim an = L ,s lim bn = K , and let c be a real number.
n→∞ n→∞

1. lim (an ± bn ) = L ± K 3. lim (an / bn ) = L/K, K ≠ 0


n→∞ n→∞

2. lim (an ⋅ bn ) = L ⋅ K 4. lim c ⋅ an = c ⋅ L


n→∞ n→∞

8.1.5 https://math.libretexts.org/@go/page/4199
Example 8.1.5: Applying properties of limits of sequences

Let the following sequences, and their limits, be given:

{ an } = {
n+1

n
2
} , and lim an = 0 ;
n→∞

n
{ bn } = {(1 +
1

n
) } , and lim bn = e ; and
n→∞

{ cn } = {n ⋅ sin(5/n)} , and lim cn = 5 .


n→∞

Evaluate the following limits.


1. lim (an + bn ) 2. lim (bn ⋅ cn ) 3. lim (1000 ⋅ an )
n→∞ n→∞ n→∞

Solution
We will use Theorem 57 to answer each of these.
1. Since lim an = 0 and lim bn = e , we conclude that lim (an + bn ) = 0 + e = e. So even though we are adding
n→∞ n→∞ n→∞

something to each term of the sequence b , we are adding something so small that the final limit is the same as before.
n

2. Since lim b = e and lim c = 5 , we conclude that lim (b ⋅ c ) = e ⋅ 5 = 5e.


n n n n
n→∞ n→∞ n→∞

3. Since lim an = 0 , we have lim 1000 an = 1000 ⋅ 0 = 0 . It does not matter that we multiply each term by 1000; the
n→∞ n→∞

sequence still approaches 0. (It just takes longer to get close to 0.)

There is more to learn about sequences than just their limits. We will also study their range and the relationships terms have with
the terms that follow. We start with some definitions describing properties of the range.

Definition 29 Bounded and unbounded sequences


A sequence {a n} is said to be bounded if there exists real numbers m and M such that m < a n <M for all n in N.
A sequence {a n} is said to be unbounded if it is not bounded.
A sequence {a } is said to be bounded above if there exists an M such that
n an < M for all n in N; it is bounded below if
there exists an m such that m < a for all n in N. n

It follows from this definition that an unbounded sequence may be bounded above or bounded below; a sequence that is both
bounded above and below is simply a bounded sequence.

Example 8.1.6: Determining boundedness of sequences

Determine the boundedness of the following sequences.


1. \(\{a_n\} = \left\{\frac1n\right\}\
2. {a } = {2 }
n
n

Solution
1. The terms of this sequence are always positive but are decreasing, so we have 0 < a < 2 for all n . Thus this sequence is
n

bounded. Figure 8.4(a) illustrates this.


2. The terms of this sequence obviously grow without bound. However, it is also true that these terms are all positive, meaning
0 < a . Thus we can say the sequence is unbounded, but also bounded below. Figure 8.4(b) illustrates this.
n

8.1.6 https://math.libretexts.org/@go/page/4199
Figure 8.4: A plot of a
n = 1/n and an = 2
n
from Example 8.1.6.

The previous example produces some interesting concepts. First, we can recognize that the sequence {1/n} converges to 0. This
says, informally, that "most'' of the terms of the sequence are "really close'' to 0. This implies that the sequence is bounded, using
the following logic. First, "most'' terms are near 0, so we could find some sort of bound on these terms (using Definition 28, the
bound is ϵ). That leaves a "few'' terms that are not near 0 (i.e., a finite number of terms). A finite list of numbers is always bounded.
This logic implies that if a sequence converges, it must be bounded. This is indeed true, as stated by the following theorem.

THEOREM 58 CONVERGENT SEQUENCES ARE BOUNDED

Let {a n} be a convergent sequence. Then {a n} is bounded.

In Example 8.1.5 we saw the sequence { bn } = { (1 + 1/n)


n
} , where it was stated that lim bn = e . (Note that this is simply
n→∞

restating part of Theorem 5.) Even though it may be difficult to intuitively grasp the behavior of this sequence, we know
immediately that it is bounded.
Another interesting concept to come out of Example 8.1.6 again involves the sequence {1/n}. We stated, without proof, that the
terms of the sequence were decreasing. That is, that a <a
n+1 for all n . (This is easy to show. Clearly n < n + 1 . Taking
n

reciprocals flips the inequality: 1/n > 1/(n + 1) . This is the same as \(a_n > a_{n+1}$.) Sequences that either steadily increase or
decrease are important, so we give this property a name.

Definition 30 MONOTONIC SEQUENCES


1. A sequence {a n} is monotonically increasing if a n ≤ an+1 for all n , i.e.,
a1 ≤ a2 ≤ a3 ≤ ⋯ an ≤ an+1 ⋯ (8.1.6)

2. A sequence {a n} is monotonically decreasing if a n ≥ an+1 for all n , i.e.,


a1 ≥ a2 ≥ a3 ≥ ⋯ an ≥ an+1 ⋯ (8.1.7)

3. A sequence is monotonic if it is monotonically increasing or monotonically decreasing.


NOTE: It is sometimes useful to call a monotonically increasing sequence strictly increasing if a < a n n+1 for all n ; i.e, we
remove the possibility that subsequent terms are equal. A similar statement holds for strictly decreasing.

8.1.7 https://math.libretexts.org/@go/page/4199
Example 8.1.7: Determining monotonicity

Determine the monotonicity of the following sequences.


2
n+1 n −9
1.{ an } = { } 3.{ an } = { }
2
n n − 10n + 26

2 2
n +1 n
2.{ an } = { } 4.{ an } = { }
n+1 n!

Solution
In each of the following, we will examine an+1− a . If a
n n+1 − an > 0 , we conclude that a < a
n n+1 and hence the sequence
is increasing. If a − a < 0 , we conclude that a > a
n+1 n n n+1 and the sequence is decreasing. Of course, a sequence need not
be monotonic and perhaps neither of the above will apply.
We also give a scatter plot of each sequence. These are useful as they suggest a pattern of monotonicity, but analytic work
should be done to confirm a graphical trend.

Figure 8.5: Plots of sequences in Example 8.1.7.


1. n+2 n+1
an+1 − an = − (8.1.8)
n+1 n
2
(n + 2)(n) − (n + 1)
= (8.1.9)
(n + 1)n

−1
= (8.1.10)
n(n + 1)

<0 for all n. (8.1.11)

8.1.8 https://math.libretexts.org/@go/page/4199
Since an+1 − an < 0 for all n , we conclude that the sequence is decreasing.

2. (n + 1 )
2
+1 n
2
+1
an+1 − an = − (8.1.12)
n+2 n+1
2 2
((n + 1 ) + 1)(n + 1) − (n + 1)(n + 2)
= (8.1.13)
(n + 1)(n + 2)
2
n + 4n + 1
= (8.1.14)
(n + 1)(n + 2)

>0 for all n. (8.1.15)

Since a n+1 − a > 0 for all n , we conclude the sequence is increasing.


n

3. We can clearly see in Figure 8.5 (c), where the sequence is plotted, that it is not monotonic. However, it does seem that after
the first 4 terms it is decreasing. To understand why, perform the same analysis as done before:
2 2
(n + 1 ) −9 n −9
an+1 − an = − (8.1.16)
2 2
(n + 1 ) − 10(n + 1) + 26 n − 10n + 26

2 2
n + 2n − 8 n −9
= − (8.1.17)
2 2
n − 8n + 17 n − 10n + 26
2 2 2 2
(n + 2n − 8)(n − 10n + 26) − (n − 9)(n − 8n + 17)
= (8.1.18)
2 2
(n − 8n + 17)(n − 10n + 26)

2
−10 n + 60n − 55
= . (8.1.19)
2 2
(n − 8n + 17)(n − 10n + 26)

We want to know when this is greater than, or less than, 0. The denominator is always positive, therefore we are only
concerned with the numerator. Using the quadratic formula, we can determine that −10n + 60n − 55 = 0 when 2

n ≈ 1.13, 4.87. So for n < 1.13 , the sequence is decreasing. Since we are only dealing with the natural numbers, this

means that a > a .


1 2

Between 1.13 and 4.87, i.e., for n = 2 , 3 and 4, we have that a >a and the sequence is increasing. (That is, when
n+1 n

n = 2 , 3 and 4, the numerator −10 n + 60n + 55 from the fraction above is > 0 .)
2

When n > 4.87, i.e, for n ≥ 5 , we have that −10n 2


+ 60n + 55 < 0 , hence a n+1 − an < 0 , so the sequence is
decreasing.

In short, the sequence is simply not monotonic. However, it is useful to note that for n ≥ 5 , the sequence is monotonically
decreasing.
4. Again, the plot in Figure 8.6 shows that the sequence is not monotonic, but it suggests that it is monotonically decreasing
after the first term. We perform the usual analysis to confirm this.
2 2
(n + 1) n
an+1 − an = − (8.1.20)
(n + 1)! n!

2 2
(n + 1 ) − n (n + 1)
= (8.1.21)
(n + 1)!
3
−n + 2n + 1
= (8.1.22)
(n + 1)!

When n = 1 , the above expression is > 0 ; for n ≥ 2 , the above expression is < 0 . Thus this sequence is not monotonic,
but it is monotonically decreasing after the first term.

8.1.9 https://math.libretexts.org/@go/page/4199
Figure 8.6: A plot of a
n
2
= n /n! in Example 8.1.7.

Knowing that a sequence is monotonic can be useful. In particular, if we know that a sequence is bounded and monotonic, we can
conclude it converges! Consider, for example, a sequence that is monotonically decreasing and is bounded below. We know the
sequence is always getting smaller, but that there is a bound to how small it can become. This is enough to prove that the sequence
will converge, as stated in the following theorem.

THEOREM 59 BOUNDED MONOTONIC SEQUENCES ARE CONVERGENT


1. Let {a n} be a bounded, monotonic sequence. Then {a n} converges; i.e., lim an exists.
n→∞

2. Let {a n} be a monotonically increasing sequence that is bounded above. Then {a n} converges.


3. Let {a n} be a monotonically decreasing sequence that is bounded below. Then {a n} converges.

Consider once again the sequence {a } = {1/n}. It is easy to show it is monotonically decreasing and that it is always positive
n

(i.e., bounded below by 0). Therefore we can conclude by Theorem 59 that the sequence converges. We already knew this by other
means, but in the following section this theorem will become very useful.
Sequences are a great source of mathematical inquiry. The On-Line Encyclopedia of Integer Sequences http://oeis.org contains
thousands of sequences and their formulae. (As of this writing, there are 257,537 sequences in the database.) Perusing this database
quickly demonstrates that a single sequence can represent several different ``real life'' phenomena.
Interesting as this is, our interest actually lies elsewhere. We are more interested in the sum of a sequence. That is, given a sequence
{ a }, we are very interested in a + a + a + ⋯ . Of course, one might immediately counter with "Doesn't this just add up to
n 1 2 3

`infinity'?'' Many times, yes, but there are many important cases where the answer is no. This is the topic of series, which we begin
to investigate in the next section.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 8.1: Sequences is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et al. via
source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

8.1.10 https://math.libretexts.org/@go/page/4199
8.2: Infinite Series
Given the sequence {a n}
n
= {1/ 2 } = 1/2, 1/4, 1/8, … , consider the following sums:
a1 = 1/2 = 1/2

a1 + a2 = 1/2 + 1/4 = 3/4


(8.2.1)
a1 + a2 + a3 = 1/2 + 1/4 + 1/8 = 7/8

a1 + a2 + a3 + a4 = 1/2 + 1/4 + 1/8 + 1/16 = 15/16

In general, we can show that


n
2 −1 1
a1 + a2 + a3 + ⋯ + an = n
=1− n
. (8.2.2)
2 2

Let S be the sum of the first n terms of the sequence {1/2 }. From the above, we see that S
n
n
1 = 1/2 ,S
2 = 3/4 , etc. Our formula
at the end shows that S = 1 − 1/2 .
n
n

Now consider the following limit:


n
lim Sn = lim (1 − 1/ 2 ) = 1. (8.2.3)
n→∞ n→∞

This limit can be interpreted as saying something amazing: the sum of all the terms of the sequence {1/2 } is 1.} This example n

illustrates some interesting concepts that we explore in this section. We begin this exploration with some definitions.

Definition 31: Infinite Series, n th


Partial Sums, Convergence, Divergence

Let {a n} be a sequence.

1. The sum ∑ an is an infinite series (or, simply series).


n=1
n

2. Let S n = ∑ ai ; the sequence {S n} is the sequence of n th


partial sums of {a n} .
i=1
∞ ∞

3. If the sequence {S n} converges to L, we say the series ∑ an converges to L, and we write ∑ an = L .


n=1 n=1

4. If the sequence {S n} diverges, the series ∑ an diverges.


n=1

∞ ∞

Using our new terminology, we can state that the series ∑ 1/ 2


n
converges, and ∑ 1/ 2
n
= 1.
n=1 n=1

We will explore a variety of series in this section. We start with two series that diverge, showing how we might discern divergence.

Example 8.2.1: Showing series diverge


1. Let {a n} = {n }
2
. Show ∑ an diverges.
n=1

2. Let {b n} = {(−1 )
n+1
} . Show ∑ bn diverges.
n=1

Solution
1. Consider S , the n
n
th
partial sum.
Sn = a1 + a2 + a3 + ⋯ + an
2 2 2 2
=1 +2 +3 ⋯ +n .

By Theorem 37, this is


n(n + 1)(2n + 1)
= . (8.2.4)
6

8.2.1 https://math.libretexts.org/@go/page/4200
∞ ∞

Since lim Sn = ∞ , we conclude that the series ∑ n


2
diverges. It is instructive to write ∑ n
2
=∞ for this tells us how
n→∞
n=1 n=1

the series diverges: it grows without bound.

A scatter plot of the sequences {a } and {S } is given in Figure 8.7(a). The terms of {a } are growing, so the terms of the
n n n

partial sums {S } are growing even faster, illustrating that the series diverges.
n

2. The sequence {b } starts with 1, −1, 1, −1, …. Consider some of the partial sums S of {b }:
n n n

S1 = 1

S2 = 0

S3 = 1

S4 = 0

1 n is odd
This pattern repeats; we find that S n ={
0 n is even

As {S n} oscillates, repeating 1, 0, 1, 0, …, we conclude that lim Sn does not exist, hence ∑ (−1 )
n+1
diverges.
n→∞
n=1

A scatter plot of the sequence {b } and the partial sums {S


n n} is given in Figure 8.7(b). When n is odd, b n = Sn so the
marks for b are drawn oversized to show they coincide.
n

Figure 8.7: Scatter plots relating to Example 8.2.1.

While it is important to recognize when a series diverges, we are generally more interested in the series that converge. In this
section we will demonstrate a few general techniques for determining convergence; later sections will delve deeper into this topic.

Geometric Series
One important type of series is a geometric series.

8.2.2 https://math.libretexts.org/@go/page/4200
Definition 32: geometric series
A geometric series is a series of the form

n 2 3 n
∑r = 1 +r+r +r +⋯ +r +⋯ (8.2.5)

n=0

Note that the index starts at n = 0 , not n = 1 .

We started this section with a geometric series, although we dropped the first term of 1. One reason geometric series are important
is that they have nice convergence properties.

theorem 60: convergence of geometric series


Consider the geometric series ∑ r


n
.
n=0

n+1
1−r
1. The n partial sum is: S =
th
n . 1−r

2. The series converges if, and only if, |r| < 1 . When |r| < 1 ,

n
1
∑r = . (8.2.6)
1 −r
n=0

According to Theorem 60, the series


∞ ∞ 2
1 1 1 1
∑ = ∑( ) =1+ + +⋯ (8.2.7)
n
2 2 2 4
n=0 n=0

converges as r = 1/2 , and ∑


1
n
2
=
1−1/2
1
= 2. This concurs with our introductory example; while there we got a sum of 1, we
n=0

skipped the first term of 1.

Example 8.2.2: Exploring geometric series

Check the convergence of the following series. If the series converges, find its sum.
∞ ∞ n ∞
3 n −1 n
1. ∑ ( ) 2. ∑ ( ) 3. ∑ 3
4 2
n=2 n=0 n=0

Solution

Figure 8.8: Scatter plots relating to the series in Example 8.2.2


1. Since r = 3/4 < 1 , this series converges. By Theorem 60, we have that
∞ n
3 1
∑( ) = = 4. (8.2.8)
4 1 − 3/4
n=0

However, note the subscript of the summation in the given series: we are to start with n = 2 . Therefore we subtract off the
first two terms, giving:

8.2.3 https://math.libretexts.org/@go/page/4200
∞ n
3 3 9
∑( ) = 4 −1 − = . (8.2.9)
4 4 4
n=2

This is illustrated in Figure 8.8.


2. Since |r| = 1/2 < 1 , this series converges, and by Theorem 60,
∞ n
−1 1 2
∑( ) = = . (8.2.10)
2 1 − (−1/2) 3
n=0

The partial sums of this series are plotted in Figure 8.9(a). Note how the partial sums are not purely increasing as some of
the terms of the sequence {(−1/2) } are negative.
n

3. Since r > 1 , the series diverges. (This makes "common sense''; we expect the sum
1 + 3 + 9 + 27 + 81 + 243 + ⋯ (8.2.11)

to diverge.) This is illustrated in Figure 8.9(b).

Figure 8.9: Scatter plots relating to the series in Example 8.2.2.

p-Series
Another important type of series is the p-series.

Definition 33: p-Series, General P -Series

1. A p--series is a series of the form



1
∑ , where p > 0. (8.2.12)
p
n
n=1

2. A general p--series} is a series of the form



1
∑ , where p > 0 and a, b are real numbers. (8.2.13)
p
(an + b)
n=1

Like geometric series, one of the nice things about p--series is that they have easy to determine convergence properties.

8.2.4 https://math.libretexts.org/@go/page/4200
theorem 61: convergence of general P --Series

A general p--series ∑
1
p
will converge if, and only if, p > 1 .
(an+b)
n=1

Note: Theorem 61 assumes that an + b ≠ 0 for all n . If an + b = 0 for some n , then of course the series does not converge
regardless of p as not all of the terms of the sequence are defined.

Example 8.2.3: Determining convergence of series

Determine the convergence of the following series.


1. ∑
1

n
n=1

2. ∑
n
1
2

n=1

3. ∑
1

√n
n=1
∞ n
(−1)
4. ∑
n
n=1

5. ∑ 1
1
3
( n−5 )
n=11 2

6. ∑
2
1
n

n=1

Solution
1. This is a p--series with p = 1 . By Theorem 61, this series diverges.
This series is a famous series, called the Harmonic Series, so named because of its relationship to harmonics in the study of
music and sound.
2. This is a p--series with p = 2 . By Theorem 61, it converges. Note that the theorem does not give a formula by which we
can determine what the series converges to; we just know it converges. A famous, unexpected result is that this series
converges to π /6. 2

3. This is a p--series with p = 1/2 ; the theorem states that it diverges.


4. This is not a p--series; the definition does not allow for alternating signs. Therefore we cannot apply Theorem 61. (Another
famous result states that this series, the Alternating Harmonic Series, converges to ln 2.)
5. This is a general p--series with p = 3 , therefore it converges.
6. This is not a p--series, but a geometric series with r = 1/2 . It converges.

Later sections will provide tests by which we can determine whether or not a given series converges. This, in general, is much
easier than determining what a given series converges to. There are many cases, though, where the sum can be determined.

Example 8.2.4: Telescoping series


Evaluate the sum ∑ (


1

n

1

n+1
) .
n=1

Solution
It will help to write down some of the first few partial sums of this series.

8.2.5 https://math.libretexts.org/@go/page/4200
1 1 1
S1 = − =1−
1 2 2

1 1 1 1 1
S2 =( − ) +( − ) =1−
1 2 2 3 3

1 1 1 1 1 1 1
S3 =( − ) +( − ) +( − ) =1−
1 2 2 3 3 4 4

1 1 1 1 1 1 1 1 1
S4 =( − ) +( − ) +( − ) +( − ) =1−
1 2 2 3 3 4 4 5 5

Note how most of the terms in each partial sum are canceled out! In general, we see that S n =1−
1

n+1
. The sequence { Sn }

converges, as lim Sn = lim (1 −


1

n+1
) =1 , and so we conclude that ∑ (
1

n

1

n+1
) =1 . Partial sums of the series are
n→∞ n→∞
n=1

plotted in Figure 8.10.

Figure 8.10: Scatter plots relating to the series of Example 8.2.4.

The series in Example 8.2.4 is an example of a telescoping series. Informally, a telescoping series is one in which the partial sums
reduce to just a finite number of terms. The partial sum S did not contain n terms, but rather just two: 1 and 1/(n + 1) .
n

When possible, seek a way to write an explicit formula for the n th


partial sum S . This makes evaluating the limit
n lim Sn much
n→∞

more approachable. We do so in the next example.


Note on notation: Most of the series we encounter will start with n = 1 . For ease of notation, we will often write ∑ a instead of n

writing ∑ an .
n=1

Example 8.2.5: Evaluating series

Evaluate each of the following infinite series.


∞ ∞

1. ∑ 2
n +2n
2
2. ∑ ln(
n+1

n
)
n=1 n=1

Solution
1. We can decompose the fraction 2/(n 2
+ 2n) as
2 1 1
= − . (8.2.14)
n2 + 2n n n+2

(See Section 6.5, Partial Fraction Decomposition, to recall how this is done, if necessary.)
Expressing the terms of {S } is now more instructive:
n

8.2.6 https://math.libretexts.org/@go/page/4200
1 1
S1 = 1 − =1−
3 3

1 1 1 1 1 1
S2 = (1 − ) +( − ) =1+ − −
3 2 4 2 3 4

1 1 1 1 1 1 1 1
S3 = (1 − ) +( − ) +( − ) =1+ − −
3 2 4 3 5 2 4 5

1 1 1 1 1 1 1 1 1 1
S4 = (1 − ) +( − ) +( − ) +( − ) =1+ − −
3 2 4 3 5 4 6 2 5 6

1 1 1 1 1 1 1 1 1 1 1 1
S5 = (1 − ) +( − ) +( − ) +( − ) +( − ) =1+ − −
3 2 4 3 5 4 6 5 7 2 6 7

We again have a telescoping series. In each partial sum, most of the terms cancel and we obtain the formula
S =1+
n −
1

2

1

n+1
. Taking limits allows us to determine the convergence of the series:
n+2
1


1 1 1 3 1 3
lim Sn = lim (1 + − − ) = , so ∑ = . (8.2.15)
2
n→∞ n→∞ 2 n+1 n+2 2 n + 2n 2
n=1

This is illustrated in Figure 8.11(a).


2. We begin by writing the first few partial sums of the series:

S1 = ln(2)

3
S2 = ln(2) + ln( )
2

3 4
S3 = ln(2) + ln( ) + ln( )
2 3

3 4 5
S4 = ln(2) + ln( ) + ln( ) + ln( )
2 3 4

At first, this does not seem helpful, but recall the logarithmic identity: ln x + ln y = ln(xy). Applying this to S gives: 4

3 4 5 2 3 4 5
S4 = ln(2) + ln( ) + ln( ) + ln( ) = ln( ⋅ ⋅ ⋅ ) = ln(5). (8.2.16)
2 3 4 1 2 3 4

We can conclude that {S n} = { ln(n + 1)} . This sequence does not converge, as lim Sn = ∞ . Therefore
n→∞

∑ ln(
n+1

n
) =∞ ; the series diverges. Note in Figure 8.11(b) how the sequence of partial sums grows slowly; after 100
n=1

terms, it is not yet over 5. Graphically we may be fooled into thinking the series converges, but our analysis above shows
that it does not.

8.2.7 https://math.libretexts.org/@go/page/4200
Figure 8.11: Scatter plots relating to the series in Example 8.2.5

We are learning about a new mathematical object, the series. As done before, we apply "old'' mathematics to this new topic.

THEOREM 62 PROPERTIES OF INFINITE SERIES


∞ ∞

Let ∑ an = L, ∑ bn = K , and let c be a constant.


n=1 n=1

∞ ∞

1. Constant Multiple Rule: ∑ c ⋅ an = c ⋅ ∑ an = c ⋅ L.


n=1 n=1
∞ ∞ ∞

2. Sum/Difference Rule: ∑ (an ± bn ) = ∑ an ± ∑ bn = L ± K.


n=1 n=1 n=1

Before using this theorem, we provide a few "famous'' series.

KEY IDEA 31 IMPORTANT SERIES


1. ∑
1

n!
=e . (Note that the index starts with n = 0 .)
n=0

2

2. ∑
1

n2
=
π

6
.
n=1
∞ n+1
(−1) 2

3. ∑ 2
n
=
π

12
.
n=1
∞ n
(−1)
4. ∑
2n+1
=
π

4
.
n=0

5. ∑
1

n
diverges . (This is called the Harmonic Series.)
n=1
∞ n+1
(−1)
6. ∑
n
= ln 2 . (This is called the Alternating Harmonic Series.)
n=1

8.2.8 https://math.libretexts.org/@go/page/4200
Example 8.2.6: Evaluating series

Evaluate the given series.


n+1
∞ 2 ∞
(−1 ) (n −n )
1000 1 1 1 1
1. ∑ 3
2. ∑ 3. + + + +⋯
n n! 16 25 36 49
n=1 n=1

Solution
1. We start by using algebra to break the series apart:
∞ n+1 2 ∞ n+1 2 n+1
(−1 ) (n − n) (−1) n (−1 ) n
∑ = ∑( − )
n3 n3 n3
n=1 n=1

∞ n+1 ∞ n+1
(−1) (−1)
=∑ −∑
2
n n
n=1 n=1

2
π
= ln(2) − ≈ −0.1293.
12

This is illustrated in Figure 8.12(a).


2. This looks very similar to the series that involves e in Key Idea 31. Note, however, that the series given in this example
starts with n = 1 and not n = 0 . The first term of the series in the Key Idea is 1/0! = 1, so we will subtract this from our
result below:
∞ ∞
1000 1
∑ = 1000 ⋅ ∑
n! n!
n=1 n=1

= 1000 ⋅ (e − 1) ≈ 1718.28.

This is illustrated in Figure 8.12(b). The graph shows how this particular series converges very rapidly.

3. The denominators in each term are perfect squares; we are adding ∑


n
1
2
(note we start with n = 4 , not n = 1 ). This series
n=4

will converge. Using the formula from Key Idea 31, we have the following:
∞ 3 ∞
1 1 1
∑ =∑ +∑
2 2 2
n n n
n=1 n=1 n=4

∞ 3 ∞
1 1 1
∑ −∑ =∑
n2 n2 n2
n=1 n=1 n=4

2 ∞
π 1 1 1 1
−( + + ) =∑
2
6 1 4 9 n
n=4

2 ∞
π 49 1
− =∑
2
6 36 n
n=4


1
0.2838 ≈ ∑
2
n
n=4

8.2.9 https://math.libretexts.org/@go/page/4200
Figure 8.12: Scatter plots relating to the series in Example 8.2.6

It may take a while before one is comfortable with this statement, whose truth lies at the heart of the study of infinite series: it is
possible that the sum of an infinite list of nonzero numbers is finite. We have seen this repeatedly in this section, yet it still may
"take some getting used to.''
As one contemplates the behavior of series, a few facts become clear.
1. In order to add an infinite list of nonzero numbers and get a finite result, "most'' of those numbers must be "very near'' 0.
2. If a series diverges, it means that the sum of an infinite list of numbers is not finite (it may approach ±∞ or it may oscillate),
and:
1. The series will still diverge if the first term is removed.
2. The series will still diverge if the first 10 terms are removed.
3. The series will still diverge if the first 1, 000, 000terms are removed.
4. The series will still diverge if any finite number of terms from anywhere in the series are removed.
These concepts are very important and lie at the heart of the next two theorems.

theorem 63 n --Term Test for Convergence/Divergence


th

Consider the series ∑ an .


n=1

1. If ∑ an converges, then lim an = 0 .


n→∞
n=1

2. If lim an ≠ 0 , then ∑ an diverges.


n→∞
n=1

Note that the two statements in Theorem 63 are really the same. In order to converge, the limit of the terms of the sequence must
approach 0; if they do not, the series will not converge.
Looking back, we can apply this theorem to the series in Example 8.2.1. In that example, the th
n terms of both sequences do not
converge to 0, therefore we can quickly conclude that each series diverges.

8.2.10 https://math.libretexts.org/@go/page/4200

Important! This theorem does not state that if lim an = 0 then ∑ an converges. The standard example of this is the Harmonic
n→∞
n=1

Series, as given in Key Idea 31. The Harmonic Sequence, {1/n}, converges to 0; the Harmonic Series, ∑ 1/n , diverges.
n=1

theorem 64 infinite nature of series


The convergence or divergence remains unchanged by the addition or subtraction of any finite number of terms. That is:
1. A divergent series will remain divergent with the addition or subtraction of any finite number of terms.
2. A convergent series will remain convergent with the addition or subtraction of any finite number of terms. (Of course, the
sum will likely change.)

Consider once more the Harmonic Series ∑


1

n
which diverges; that is, the sequence of partial sums { Sn } grows (very, very
n=1

slowly) without bound. One might think that by removing the "large'' terms of the sequence that perhaps the series will converge.
This is simply not the case. For instance, the sum of the first 10 million terms of the Harmonic Series is about 16.7. Removing the
first 10 million terms from the Harmonic Series changes the n partial sums, effectively subtracting 16.7 from the sum. However,
th

a sequence that is growing without bound will still grow without bound when 16.7 is subtracted from it.
The equations below illustrate this. The first line shows the infinite sum of the Harmonic Series split into the sum of the first 10
million terms plus the sum of "everything else.'' The next equation shows us subtracting these first 10 million terms from both
sides. The final equation employs a bit of "psuedo--math'': subtracting 16.7 from "infinity'' still leaves one with "infinity.''
∞ 10,000,000 ∞
1 1 1
∑ = ∑ + ∑
n n n
n=1 n=1 n=10,000,001

∞ 10,000,000 ∞
1 1 1
∑ − ∑ = ∑
n n n
n=1 n=1 n=10,000,001

∞ − 16.7 = ∞.

This section introduced us to series and defined a few special types of series whose convergence properties are well known: we
know when a p-series or a geometric series converges or diverges. Most series that we encounter are not one of these types, but we
are still interested in knowing whether or not they converge. The next three sections introduce tests that help us determine whether
or not a given series converges.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 8.2: Infinite Series is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et al.
via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

8.2.11 https://math.libretexts.org/@go/page/4200
8.3: Integral and Comparison Tests
Knowing whether or not a series converges is very important, especially when we discusses Power Series. Theorems 60 and 61
give criteria for when Geometric and p-series converge, and Theorem 63 gives a quick test to determine if a series diverges. There
are many important series whose convergence cannot be determined by these theorems, though, so we introduce a set of tests that
allow us to handle a broad range of series. We start with the Integral Test.

Integral Test
We stated in Section 8.1 that a sequence {a n} is a function a(n) whose domain is N, the set of natural numbers. If we can extend

a(n) to R , the real numbers, and it is both positive and decreasing on [1, ∞) , then the convergence of ∑ an is the same as
n=1

∫ a(x)dx .
1

theorem 8.3.1: integral test


Let a sequence { an } be defined by an = a(n) , where a(n) is continuous, positive and decreasing on [1, ∞) . Then ∑ an
n=1

converges, if, and only if, ∫ a(x)dx converges.


1

We can demonstrate the truth of the Integral Test with two simple graphs. In Figure 8.3.1a, the height of each rectangle is
a(n) = a for n = 1, 2, …, and clearly the rectangles enclose more area than the area under y = a(x) . Therefore we can conclude
n

that

∫ a(x)dx < ∑ an . (8.3.1)

n=1
1

Figure 8.3.1 : Illustrating the truth of the Integral Test.


In Figure 8.3.1b , we draw rectangles under y = a(x) with the Right-Hand rule, starting with n =2 . This time, the area of the
∞ ∞

rectangles is less than the area under y = a(x) , so ∑ an < ∫ a(x)dx . Note how this summation starts with n = 2 ; adding a to 1

n=2 1

both sides lets us rewrite the summation starting with n = 1 :



∑ an < a1 + ∫ a(x)dx. (8.3.2)

n=1
1

Combining Equations 8.3.1 and 8.3.2, we have



∞ ∞

∑ an < a1 + ∫ a(x)dx < a1 + ∑ an . (8.3.3)

n=1 n=1
1

8.3.1 https://math.libretexts.org/@go/page/4201
Theorem 8.3.1

From Equation 8.3.3 we can make the following two statements:


∞ ∞ ∞ ∞

1. If ∑ an diverges, so does ∫ a(x)dx (because ∑ an < a1 + ∫ a(x)dx)


n=1 1 n=1 1
∞ ∞ ∞ ∞

2. If ∑ an converges, so does ∫ a(x)dx (because ∫ a(x)dx < ∑ an . )


n=1 1 1 n=1

Therefore the series and integral either both converge or both diverge.

Theorem8.3.1 allows us to extend this theorem to series where a(n) is positive and decreasing on [b, ∞) for some b > 1 .

Example 8.3.1: Using the Integral Test



ln n
Determine the convergence of ∑
2
. (The terms of the sequence { an } = {ln n/ n }
2
and the nth
partial sums are given in
n=1 n

Figure 8.3.2).

Solution
Figure 8.3.2 implies that a(n) = (ln n)/n is positive and decreasing on [2, ∞). We can determine this analytically, too. We
2

know a(n) is positive as both ln n and n are positive on [2, ∞). To determine that a(n) is decreasing, consider
2

a (n) = (1 − 2 ln n)/ n , which is negative for n ≥ 2 . Since a (n) is negative, a(n) is decreasing.
′ 3 ′

Figure 8.3.2 : Plotting the sequence and series in Example 8.3.1 .



ln x
Applying the Integral Test, we test the convergence of ∫
2
dx . Integrating this improper integral requires the use of
1
x

Integration by Parts, with u = ln x and dv = 1/x 2


dx .
∞ b

ln x ln x
∫ dx = lim ∫ dx
2 2
x b→∞ x
1 1

1 b 1

= lim − ln x +∫ dx
∣ 2
b→∞ x 1 x
1

1 1 b

= lim − ln x −

b→∞ x x 1

1 ln b
= lim 1 − − . ^pital's Rule:
Apply L'Ho
b→∞ b b

= 1.

∞ ∞
ln x ln n
Since ∫ 2
dx converges, so does ∑
2
.
1
x n=1 n


1
Theorem 61 was given without justification, stating that the general p-series ∑
p
converges if, and only if, p > 1 . In the
n=1 (an + b)

following example, we prove this to be true by applying the Integral Test.

8.3.2 https://math.libretexts.org/@go/page/4201
Example 8.3.2: Using the Integral Test to establish Theorem 61

1
Use the Integral Test to prove that ∑
p
converges if, and only if, p > 1 .
n=1 (an + b)

Solution

1
Consider the integral ∫ p
dx ; assuming p ≠ 1 ,
1
(ax + b)

∞ c

1 1
∫ dx = lim ∫ dx
p c→∞ p
(ax + b) (ax + b)
1 1

1 c
1−p ∣
= lim (ax + b )

c→∞ a(1 − p) 1

1
1−p 1−p
= lim ((ac + b ) − (a + b ) ).
c→∞ a(1 − p)

This limit converges if, and only if, p > 1 . It is easy to show that the integral also diverges in the case of p = 1 . (This result is
similar to the work preceding Key Idea 21.)

1
Therefore ∑
p
converges if, and only if, p > 1 .
n=1 (an + b)

We consider two more convergence tests in this section, both comparison tests. That is, we determine the convergence of one series
by comparing it to another series with known convergence.

Direct Comparison Test


theorem 8.3.1: direct comparison test

Let {a n} and {b n} be positive sequences where a n ≤ bn for all n ≥ N , for some N ≥1 .


∞ ∞

1. If ∑ bn converges, then ∑ an converges.


n=1 n=1
∞ ∞

2. If ∑ an diverges, then ∑ bn diverges.


n=1 n=1

Note: A sequence {a n} is a positive sequence if a n >0 for all n .


Because of Theorem 64, any theorem that relies on a positive sequence still holds true when a n >0 for all but a finite number of
values of n .

Example 8.3.3: Applying the Direct Comparison Test



1
Determine the convergence of ∑
n 2
.
n=1 3 +n

Solution
This series is neither a geometric or p-series, but seems related. We predict it will converge, so we look for a series with larger
terms that converges. (Note too that the Integral Test seems difficult to apply here.)

1 1 1
Since 3
n
<3
n
+n
2
, n
>
n 2
for all n ≥1 . The series ∑
n
is a convergent geometric series; by Theorem 66,
3 3 +n n=1
3

1

n
converges.
n=1 3 + n2

8.3.3 https://math.libretexts.org/@go/page/4201
Example 8.3.4: Applying the Direct Comparison Test

1
Determine the convergence of ∑ .
n=1
n − ln n

Solution

1
We know the Harmonic Series ∑ diverges, and it seems that the given series is closely related to it, hence we predict it will
n=1
n

diverge.
1 1
Since n ≥ n − ln n for all n ≥ 1 , ≤ for all n ≥ 1 .
n n − ln n


1
The Harmonic Series diverges, so we conclude that ∑ diverges as well.
n=1
n − ln n

The concept of direct comparison is powerful and often relatively easy to apply. Practice helps one develop the necessary intuition
to quickly pick a proper series with which to compare. However, it is easy to construct a series for which it is difficult to apply the
Direct Comparison Test.

1
Consider ∑ . It is very similar to the divergent series given in Example 8.3.5. We suspect that it also diverges, as
n=1
n + ln n

1 1
≈ for large n . However, the inequality that we naturally want to use "goes the wrong way'': since n ≤ n + ln n for
n n + ln n
1 1
all n ≥1 , ≥ for all n ≥1 . The given series has terms less than the terms of a divergent series, and we cannot
n n + ln n

conclude anything from this.


Fortunately, we can apply another test to the given series to determine its convergence.

Large Limit Comparison Test


Theorem 67: limit comparison test
Let {a n} and {b n} be positive sequences.
∞ ∞
an
1. If lim n→∞ =L , where L is a positive real number, then ∑ an and ∑ bn either both converge or both diverge.
bn n=1 n=1
∞ ∞
an
2. If lim n→∞ =0 , then if ∑ bn converges, then so does ∑ an .
bn n=1 n=1
∞ ∞
an
3. If lim n→∞ =∞ , then if ∑ bn diverges, then so does ∑ an .
bn n=1 n=1

Theorem 67 is most useful when the convergence of the series from {b n} is known and we are trying to determine the convergence
of the series from {a }. n


1
We use the Limit Comparison Test in the next example to examine the series ∑ which motivated this new test.
n=1
n + ln n

Example 8.3.5: Applying the Limit Comparison Test



1
Determine the convergence of ∑ using the Limit Comparison Test.
n=1 n + ln n

Solution
∞ ∞
1 1
We compare the terms of ∑ to the terms of the Harmonic Sequence ∑ :
n=1
n + ln n n=1
n

8.3.4 https://math.libretexts.org/@go/page/4201
1/(n + ln n) n
lim = lim
n→∞ 1/n n→∞ n + ln n

=1 ^pital's Rule).
(after applying L'Ho


1
Since the Harmonic Series diverges, we conclude that ∑ diverges as well.
n=1
n + ln n

Example 8.3.6: Applying the Limit Comparison Test



1
Determine the convergence of ∑
n 2
n=1 3 −n

Solution
This series is similar to the one in Example 8.3.3, but now we are considering "3 n 2
−n '' instead of "3 n
+n
2
.'' This difference
makes applying the Direct Comparison Test difficult.

1
Instead, we use the Limit Comparison Test and compare with the series ∑
n
:
n=1 3

n 2 n
1/(3 −n ) 3
lim = lim
n n 2
n→∞ 1/3 n→∞ 3 −n

=1 (after applying L'Ho


^pital's Rule twice).

∞ ∞
1 1
We know ∑
n
is a convergent geometric series, hence ∑
n 2
converges as well.
n=1
3 n=1 3 −n

As mentioned before, practice helps one develop the intuition to quickly choose a series with which to compare. A general rule of
thumb is to pick a series based on the dominant term in the expression of {a }. It is also helpful to note that factorials dominate
n

exponentials, which dominate algebraic functions (e.g., polynomials), which dominate logarithms. In the previous example, the

1 1
dominant term of n 2
was 3 , so we compared the series to
n

n
. It is hard to apply the Limit Comparison Test to series
3 −n n=1
3

containing factorials, though, as we have not learned how to apply L'Ho^pital's Rule to n!.

Example 8.3.7: Applying the Limit Comparison Test




√n + 3
Determine the convergence of ∑
2
.
n=1 n −n+1

Solution
We naively attempt to apply the rule of thumb given above and note that the dominant term in the expression of the series is

1
1/n
2
. Knowing that ∑
2
converges, we attempt to apply the Limit Comparison Test:
n=1 n

− 2 2 −
(√n + 3)/(n − n + 1) n (√n + 3)
lim = lim
n→∞ 2 n→∞ 2
1/n n −n+1

=∞ ^pital's Rule).
(Apply L'Ho

Theorem 67 part (3) only applies when ∑ bn diverges; in our case, it converges. Ultimately, our test has not revealed anything
n=1

about the convergence of our series.


The problem is that we chose a poor series with which to compare. Since the numerator and denominator of the terms of the
series are both algebraic functions, we should have compared our series to the dominant term of the numerator divided by the
dominant term of the denominator.
The dominant term of the numerator is n 1/2
and the dominant term of the denominator is n
2
. Thus we should compare the
terms of the given series to n /n = 1/n1/2 2 3/2
:

8.3.5 https://math.libretexts.org/@go/page/4201
− 2 3/2 −
(√n + 3)/(n − n + 1) n (√n + 3)
lim = lim
3/2
n→∞ 1/n n→∞ n2 − n + 1

=1 (Apply L'Ho
^pital's Rule).

∞ ∞

1 √n + 3
Since the p-series ∑
3/2
converges, we conclude that ∑
2
converges as well.
n=1 n n=1 n −n+1

We mentioned earlier that the Integral Test did not work well with series containing factorial terms. The next section introduces the
Ratio Test, which does handle such series well. We also introduce the Root Test, which is good for series where each term is
raised to a power.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 8.3: Integral and Comparison Tests is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by
Gregory Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available
upon request.

8.3.6 https://math.libretexts.org/@go/page/4201
8.4: Ratio and Root Tests

The n
th
--Term Test of Theorem 63 states that in order for a series ∑ an to converge, lim an = 0 . That is, the terms of { an }
n→∞
n=1

must get very small. Not only must the terms approach 0, they must approach 0 "fast enough'': while lim 1/n = 0 , the Harmonic
n→∞

1
Series ∑ diverges as the terms of {1/n} do not approach 0 "fast enough.''
n=1 n

The comparison tests of the previous section determine convergence by comparing terms of a series to terms of another series
whose convergence is known. This section introduces the Ratio and Root Tests, which determine convergence by analyzing the
terms of a series to see if they approach 0 "fast enough.''

Ratio Test
theorem 68: ratio test
an+1
Let {a n} be a positive sequence where lim =L .
n→∞ an

1. If L < 1 , then ∑ an converges.


n=1

2. If L > 1 or L = ∞ , then ∑ an diverges.


n=1

3. If L = 1 , the Ratio Test is inconclusive.

Theorem 64 allows us to apply the Ratio Test to series where {a n} is positive for all but a finite number of terms.
an+1
The principle of the Ratio Test is this: if lim =L <1 , then for large n , each term of {a n} is significantly smaller than its
n→∞ an

previous term which is enough to ensure convergence.

Example 8.4.1: Applying the Ratio Test

Use the Ratio Test to determine the convergence of the following series:
∞ n
2
1. ∑ .
n=1
n!
∞ n
3
2. ∑
3
n=1 n

1
3. ∑
2
.
n=1 n +1

Solution
∞ n
2
1. ∑ :
n=1
n!

n+1 n+1
2 /(n + 1)! 2 n!
lim = lim
n n
n→∞ 2 /n! n→∞ 2 (n + 1)!

2
= lim
n→∞ n+1

= 0.

∞ n
2
Since the limit is 0 < 1 , by the Ratio Test ∑ converges.
n=1
n!
∞ n
3
2. ∑
3
:
n=1 n

8.4.1 https://math.libretexts.org/@go/page/4202
n+1 3 n+1 3
3 /(n + 1 ) 3 n
lim = lim
n 3 n 3
n→∞ 3 /n n→∞ 3 (n + 1 )
3
3n
= lim
n→∞ 3
(n + 1)

= 3.

∞ n
3
Since the limit is 3 > 1 , by the Ratio Test ∑
3
diverges.
n=1 n

1
3. ∑
2
:
n=1 n +1

2 2
1/((n + 1 ) + 1) n +1
lim = lim
n→∞ 2 n→∞ 2
1/(n + 1) (n + 1 ) +1

= 1.

Since the limit is 1, the Ratio Test is inconclusive. We can easily show this series converges using the Direct or Limit

1
Comparison Tests, with each comparing to the series ∑
2
.
n=1 n

The Ratio Test is not effective when the terms of a series only contain algebraic functions (e.g., polynomials). It is most effective
when the terms contain some factorials or exponentials. The previous example also reinforces our developing intuition: factorials
dominate exponentials, which dominate algebraic functions, which dominate logarithmic functions. In Part 1 of the example, the
factorial in the denominator dominated the exponential in the numerator, causing the series to converge. In Part 2, the exponential
in the numerator dominated the algebraic function in the denominator, causing the series to diverge.
While we have used factorials in previous sections, we have not explored them closely and one is likely to not yet have a strong
intuitive sense for how they behave. The following example gives more practice with factorials.

Example 8.4.2: Applying the Ratio Test



n!n!
Determine the convergence of ∑ .
n=1 (2n)!

Solution
Before we begin, be sure to note the difference between (2n)! and 2n! . When n =4 , the former is
8! = 8 ⋅ 7 ⋅ … ⋅ 2 ⋅ 1 = 40, 320 , whereas the latter is 2(4 ⋅ 3 ⋅ 2 ⋅ 1) = 48 .

Applying the Ratio Test:

(n + 1)!(n + 1)!/(2(n + 1))! (n + 1)!(n + 1)!(2n)!


lim = lim
n→∞ n!n!/(2n)! n→∞ n!n!(2n + 2)!

Noting that (2n + 2)! = (2n + 2) ⋅ (2n + 1) ⋅ (2n)!, we have

(n + 1)(n + 1)
= lim
n→∞ (2n + 2)(2n + 1)

= 1/4.


n!n!
Since the limit is 1/4 < 1 , by the Ratio Test we conclude ∑ converges.
n=1 (2n)!

Root Test
The final test we introduce is the Root Test, which works particularly well on series where each term is raised to a power, and does
not work well with terms containing factorials.

8.4.2 https://math.libretexts.org/@go/page/4202
theorem 69: root test

Let {a n} be a positive sequence and let lim (an )


1/n
=L .
n→∞

1. If L < 1 , then ∑ an converges.


n=1

2. If L > 1 or L = ∞ , then ∑ an diverges.


n=1

3. If L = 1 , the Root Test is inconclusive.

Example 8.4.3: Applying the Root Test

Determine the convergence of the following series using the Root Test:
n 4 n
∞ ∞ ∞
3n + 1 n 2
1. ∑ ( ) 2. ∑
n
3. ∑ .
n=1 5n − 2 n=1 (ln n) n=1 n2

Solution
n 1/n
3n + 1 3n + 1 3
1. lim ( ( ) ) = lim = .
n→∞ 5n − 2 n→∞ 5n − 2 5

Since the limit is less than 1, we conclude the series converges. Note: it is difficult to apply the Ratio Test to this series.
1/n 4
4 1/n
n (n )
2. lim (
n
) = lim .
n→∞ (ln n) n→∞ ln n

As n grows, the numerator approaches 1 (apply L'H\^opital's Rule) and the denominator grows to infinity. Thus
4
1/n
(n )
lim = 0. (8.4.1)
n→∞ ln n

Since the limit is less than 1, we conclude the series converges.


n 1/n
2 2
3. lim (
2
) = lim
2
=2 .
n→∞ n n→∞ 1/n
(n )

Since this is greater than 1, we conclude the series diverges.

Each of the tests we have encountered so far has required that we analyze series from positive sequences. The next section relaxes
this restriction by considering alternating series, where the underlying sequence has terms that alternate between being positive and
negative.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 8.4: Ratio and Root Tests is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory
Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

8.4.3 https://math.libretexts.org/@go/page/4202
8.5: Alternating Series and Absolute Convergence
All of the series convergence tests we have used require that the underlying sequence {a } be a positive sequence. (We can relax n

this with Theorem 64 and state that there must be an N > 0 such that a > 0 for all n > N ; that is, {a } is positive for all but a
n n

finite number of values of n .)


In this section we explore series whose summation includes negative terms. We start with a very specific form of series, where the
terms of the summation alternate between being positive and negative.

Definition 34: alternating series


Let {a n} be a positive sequence. An alternating series is a series of either the form
∞ ∞

n n+1
∑(−1 ) an or ∑(−1 ) an . (8.5.1)

n=1 n=1

Recall the terms of Harmonic Series come from the Harmonic Sequence { an } = {1/n} . An important alternating series is the
Alternating Harmonic Series:

1 1 1 1 1 1
n+1
∑(−1 ) =1− + − + − +⋯ (8.5.2)
n 2 3 4 5 6
n=1

Geometric Series can also be alternating series when r < 0 . For instance, if r = −1/2 , the geometric series is
∞ n
−1 1 1 1 1 1
∑( ) =1− + − + − +⋯ (8.5.3)
2 2 4 8 16 32
n=0


1
Theorem 60 states that geometric series converge when |r| < 1 and gives the sum: ∑ r
n
= . When r = −1/2 as above, we
n=0 1 −r

find
∞ n
−1 1 1 2
∑( ) = = = . (8.5.4)
2 1 − (−1/2) 3/2 3
n=0

A powerful convergence theorem exists for other alternating series that meet a few conditions.

theorem 70: alternating series test


Let {a n} be a positive, decreasing sequence where lim an = 0 . Then
n→∞

∞ ∞

n n+1
∑(−1 ) an and ∑(−1 ) an (8.5.5)

n=1 n=1

converge.

The basic idea behind Theorem 70 is illustrated in Figure 8.5.1 . A positive, decreasing sequence { an } is shown along with the
partial sums
n

i+1 n+1
Sn = ∑(−1 ) ai = a1 − a2 + a3 − a4 + ⋯ + (−1 ) an . (8.5.6)

i=1

Because {a } is decreasing, the amount by which S bounces up/down decreases. Moreover, the odd terms of S form a
n n n

decreasing, bounded sequence, while the even terms of S form an increasing, bounded sequence. Since bounded, monotonic
n

sequences converge (see Theorem 59) and the terms of {a } approach 0, one can show the odd and even terms of S converge to
n n

the same common limit L, the sum of the series.

8.5.1 https://math.libretexts.org/@go/page/4203
Figure 8.5.1 : Illustrating the convergence with the Alternating Series Test.

Example 8.5.1: Applying the Alternating Series Test

Determine if the Alternating Series Test applies to each of the following series.

1
1. ∑ (−1 )
n+1

n=1
n

ln n
2. ∑ (−1 )
n

n=1 n

| sin n|
3. ∑ (−1 )
n+1

2
n=1 n

Solution
1. This is the Alternating Harmonic Series as seen previously. The underlying sequence is {a } = {1/n}, which is positive,
n

decreasing, and approaches 0 as n → ∞ . Therefore we can apply the Alternating Series Test and conclude this series
converges.

1
While the test does not state what the series converges to, we will see later that ∑ (−1 )
n+1
= ln 2.
n=1 n

2. The underlying sequence is {a } = {ln n/n}. This is positive and approaches 0 as n → ∞ (use L'Hopital's Rule).
n

However, the sequence is not decreasing for all n . It is straightforward to compute a = 0 , a ≈ 0.347 , a ≈ 0.366 , and
1 2 3

a ≈ 0.347 : the sequence is increasing for at least the first 3 terms.


4

We do not immediately conclude that we cannot apply the Alternating Series Test. Rather, consider the long--term behavior
of {a }. Treating a = a(n) as a continuous function of n defined on [1, ∞), we can take its derivative:
n n


1 − ln n
a (n) = . (8.5.7)
2
n

The derivative is negative for all n ≥ 3 (actually, for all n > e ), meaning a(n) = a is decreasing on [3, ∞). We can
n

ln n
apply the Alternating Series Test to the series when we start with n = 3 and conclude that ∑ (−1 )
n
converges;
n=3 n

adding the terms with n = 1 and n = 2 do not change the convergence (i.e., we apply Theorem 64).

The important lesson here is that as before, if a series fails to meet the criteria of the Alternating Series Test on only a finite
number of terms, we can still apply the test.
3. The underlying sequence is {a } = | sin n|/n. This sequence is positive and approaches 0 as n → ∞ . However, it is not a
n

decreasing sequence; the value of | sin n| oscillates between 0 and 1 as n → ∞ . We cannot remove a finite number of
terms to make {a } decreasing, therefore we cannot apply the Alternating Series Test.
n

Keep in mind that this does not mean we conclude the series diverges; in fact, it does converge. We are just unable to
conclude this based on Theorem 70.

Key Idea 31 gives the sum of some important series. Two of these are

8.5.2 https://math.libretexts.org/@go/page/4203
∞ 2 ∞ n+1 2
1 π (−1) π
∑ = ≈ 1.64493 and ∑ = ≈ 0.82247. (8.5.8)
2 2
n 6 n 12
n=1 n=1

These two series converge to their sums at different rates. To be accurate to two places after the decimal, we need 202 terms of the
first series though only 13 of the second. To get 3 places of accuracy, we need 1069 terms of the first series though only 33 of the
second. Why is it that the second series converges so much faster than the first?
While there are many factors involved when studying rates of convergence, the alternating structure of an alternating series gives
us a powerful tool when approximating the sum of a convergent series.

theorem 71: the alternating series approximation theorem


Let {a n} be a sequence that satisfies the hypotheses of the Alternating Series Test, and let Sn and L be the n
th
partial sums
∞ ∞

and sum, respectively, of either ∑ (−1 ) an


n
or ∑ (−1 )
n+1
an . Then
n=1 n=1

1. |S − L| < a
n , and n+1

2. L is between S and S n n+1 .

Part 1 of Theorem 71 states that the n th


partial sum of a convergent alternating series will be within a n+1 of its total sum. Consider
∞ n+1
(−1)
the alternating series we looked at before the statement of the theorem, ∑ . Since a 14 = 1/ 14
2
≈ 0.0051 , we know that
n=1 n2

S13 is within 0.0051 of the total sum.


Moreover, Part 2 of the theorem states that since S ≈ 0.8252 and S ≈ 0.8201, we know the sum L lies between
13 14 0.8201 and
0.8252. One use of this is the knowledge that S is accurate to two places after the decimal.
14


ln n
Some alternating series converge slowly. In Example 8.5.1 we determined the series ∑ (−1 )
n+1
converged. With n = 1001,
n=1
n

we find ln n/n ≈ 0.0069, meaning that S ≈ 0.1633 is accurate to one, maybe two, places after the decimal. Since
1000

S1001 ≈ 0.1564 , we know the sum L is 0.1564 ≤ L ≤ 0.1633.

Example 8.5.2: Approximating the sum of convergent alternating series

Approximate the sum of the following series, accurate to within 0.001.


∞ ∞
1 ln n
1. ∑ (−1) n+1

3
2. ∑ (−1 )
n+1
.
n=1 n n=1
n

Solution
1. Using Theorem 71, we want to find n where 1/n 3
< 0.001 :
1 1
≤ 0.001 =
3
n 1000
3
n ≥ 1000
3 −−−−
n ≥ √1000

n ≥ 10.

Let L be the sum of this series. By Part 1 of the theorem, |S 9 − L| < a10 = 1/1000 . We can compute S 9 = 0.902116 ,
which our theorem states is within 0.001 of the total sum.

We can use Part 2 of the theorem to obtain an even more accurate result. As we know the 10 term of the series is th

−1/1000, we can easily compute S = 0.901116. Part 2 of the theorem states that L is between S and S
10 , so 9 10

0.901116 < L < 0.902116 .


2. We want to find n where ln(n)/n < 0.001. We start by solving ln(n)/n = 0.001 for n . This cannot be solved
algebraically, so we will use Newton's Method to approximate a solution.

Let f (x) = ln(x)/x − 0.001; we want to know where f (x) = 0 . We make a guess that x must be "large,'' so our initial

8.5.3 https://math.libretexts.org/@go/page/4203
guess will be x = 1000. Recall how Newton's Method works: given an approximate solution x , our next approximation
1 n

x is given by
n+1

f (xn )
xn+1 = xn − . (8.5.9)

f (xn )

We find f ′
(x) = (1 − ln(x))/ x
2
. This gives
ln(1000)/1000 − 0.001
x2 = 1000 −
2
(1 − ln(1000))/1000

= 2000.

Using a computer, we find that Newton's Method seems to converge to a solution x = 9118.01 after 8 iterations. Taking the
next integer higher, we have n = 9119, where ln(9119)/9119 = 0.000999903 < 0.001 .

Again using a computer, we find S 9118 = −0.160369 . Part 1 of the theorem states that this is within 0.001 of the actual
sum L. Already knowing the 9,119 th
term, we can compute S = −0.159369, meaning −0.159369 < L < −0.160369
9119 .
Notice how the first series converged quite quickly, where we needed only 10 terms to reach the desired accuracy, whereas the
second series took over 9,000 terms.


1
One of the famous results of mathematics is that the Harmonic Series, ∑ diverges, yet the Alternating Harmonic Series,
n=1 n

1
∑ (−1 )
n+1
, converges. The notion that alternating the signs of the terms in a series can make a series converge leads us to the
n=1
n

following definitions.

Definition 35: absolute and conditional convergence


∞ ∞

1. A series ∑ an converges absolutely if ∑ | an | converges.


n=1 n=1
∞ ∞ ∞

2. A series ∑ an converges conditionally if ∑ an converges but ∑ | an | diverges.


n=1 n=1 n=1

Thus we say the Alternating Harmonic Series converges conditionally.

Example 8.5.3: Determining absolute and conditional convergence.

Determine if the following series converge absolutely, conditionally, or diverge.


∞ ∞ 2 ∞
n+3 n + 2n + 5 3n − 3
1. ∑ (−1) n

2
2. ∑ (−1 )
n
n
3. ∑ (−1 )
n

n=1 n + 2n + 5 n=1 2 n=3 5n − 10

Solution
1. We can show the series
∞ ∞
∣ n
n+3 ∣ n+3
∑ ∣(−1 ) ∣ =∑ (8.5.10)
∣ n2 + 2n + 5 ∣ n2 + 2n + 5
n=1 n=1

diverges using the Limit Comparison Test, comparing with 1/n.


n+3
The series ∑ (−1 )
n

2
converges using the Alternating Series Test; we conclude it converges conditionally.
n=1 n + 2n + 5

8.5.4 https://math.libretexts.org/@go/page/4203
2. We can show the series
∞ 2 2 ∞
∣ n + 2n + 5 ∣ n + 2n + 5
n
∑ ∣(−1 ) ∣ =∑ (8.5.11)
n n
∣ 2 ∣ 2
n=1 n=1

converges using the Ratio Test.

∞ 2
n + 2n + 5
Therefore we conclude ∑ (−1 )
n
n
converges absolutely.
n=1
2

3. The series
∞ ∞
∣ 3n − 3 ∣ 3n − 3
n
∑ ∣(−1 ) ∣ =∑ (8.5.12)
∣ 5n − 10 ∣ 5n − 10
n=3 n=3

diverges using the n th


Term Test, so it does not converge absolutely.


3n − 3
The series ∑ (−1 )
n
fails the conditions of the Alternating Series Test as (3n − 3)/(5n − 10) does not approach
n=3 5n − 10

0 as n → ∞ . We can state further that this series diverges; as n → ∞ , the series effectively adds and subtracts 3/5 over
and over. This causes the sequence of partial sums to oscillate and not converge.


3n − 3
Therefore the series ∑ (−1 )
n
diverges.
n=1 5n − 10

Knowing that a series converges absolutely allows us to make two important statements, given in the following theorem. The first

is that absolute convergence is "stronger'' than regular convergence. That is, just because ∑ an converges, we cannot conclude
n=1
∞ ∞

that ∑ | an | will converge, but knowing a series converges absolutely tells us that ∑ an will converge.
n=1 n=1

One reason this is important is that our convergence tests all require that the underlying sequence of terms be positive. By taking
the absolute value of the terms of a series where not all terms are positive, we are often able to apply an appropriate test and
determine absolute convergence. This, in turn, determines that the series we are given also converges.
The second statement relates to rearrangements of series. When dealing with a finite set of numbers, the sum of the numbers does
not depend on the order which they are added. (So 1 + 2 + 3 = 3 + 1 + 2 .) One may be surprised to find out that when dealing
with an infinite set of numbers, the same statement does not always hold true: some infinite lists of numbers may be rearranged in
different orders to achieve different sums. The theorem states that the terms of an absolutely convergent series can be rearranged in
any way without affecting the sum.

theorem 72: absolute convergence theorem


Let ∑ an be a series that converges absolutely.


n=1

1. ∑ an converges.
n=1

2. Let {b n} be any rearrangement of the sequence {a n} . Then


∞ ∞

∑ bn = ∑ an . (8.5.13)

n=1 n=1

In Example 8.5.3, we determined the series in part 2 converges absolutely. Theorem 72 tells us the series converges (which we
could also determine using the Alternating Series Test).
The theorem states that rearranging the terms of an absolutely convergent series does not affect its sum. This implies that perhaps
the sum of a conditionally convergent series can change based on the arrangement of terms. Indeed, it can. The Riemann

8.5.5 https://math.libretexts.org/@go/page/4203
Rearrangement Theorem (named after Bernhard Riemann) states that any conditionally convergent series can have its terms
rearranged so that the sum is any desired value, including ∞!
As an example, consider the Alternating Harmonic Series once more. We have stated that

n+1
1 1 1 1 1 1 1
∑(−1 ) =1− + − + − + ⋯ = ln 2, (8.5.14)
n 2 3 4 5 6 7
n=1

(see Key Idea 31 or Example 8.5.1).


Consider the rearrangement where every positive term is followed by two negative terms:
1 1 1 1 1 1 1 1
1− − + − − + − − ⋯ (8.5.15)
2 4 3 6 8 5 10 12

(Convince yourself that these are exactly the same numbers as appear in the Alternating Harmonic Series, just in a different order.)
Now group some terms and simplify:
1 1 1 1 1 1 1 1
(1 − )− +( − )− +( − )− +⋯ =
2 4 3 6 8 5 10 12

1 1 1 1 1 1
− + − + − +⋯ =
2 4 6 8 10 12

1 1 1 1 1 1 1
(1 − + − + − + ⋯) = ln 2.
2 2 3 4 5 6 2

By rearranging the terms of the series, we have arrived at a different sum! (One could try to argue that the Alternating Harmonic
Series does not actually converge to ln 2, because rearranging the terms of the series shouldn't change the sum. However, the
Alternating Series Test proves this series converges to L, for some number L, and if the rearrangement does not change the sum,
then L = L/2 , implying L = 0 . But the Alternating Series Approximation Theorem quickly shows that L > 0 . The only
conclusion is that the rearrangement \emph{did} change the sum.) This is an incredible result.
We end here our study of tests to determine convergence. The back cover of this text contains a table summarizing the tests that one
may find useful.
While series are worthy of study in and of themselves, our ultimate goal within calculus is the study of Power Series, which we will
consider in the next section. We will use power series to create functions where the output is the result of an infinite summation.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 8.5: Alternating Series and Absolute Convergence is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or
curated by Gregory Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history
is available upon request.

8.5.6 https://math.libretexts.org/@go/page/4203
8.6: Power Series
So far, our study of series has examined the question of "Is the sum of these infinite terms finite?,'' i.e., "Does the series converge?''
We now approach series from a different perspective: as a function. Given a value of x, we evaluate f (x) by finding the sum of a
particular series that depends on x (assuming the series converges). We start this new approach to series with a definition.

Definition 36: power series


Let {a n} be a sequence, let x be a variable, and let c be a real number.
1. The power series in x is the series

n 2 3
∑ an x = a0 + a1 x + a2 x + a3 x +… (8.6.1)

n=0

2. The power series in x centered at c is the series



n 2 3
∑ an (x − c ) = a0 + a1 (x − c) + a2 (x − c ) + a3 (x − c ) +… (8.6.2)

n=0

Example 8.6.1: Examples of power series

Write out the first five terms of the following power series:
∞ ∞ n ∞ 2n
(x+1) (x−π)
n n+1 n+1
1. ∑ x 2. ∑ (−1 ) 3. ∑ (−1 ) .
n (2n)!
n=0 n=1 n=0

Solution
1. One of the conventions we adopt is that x 0
=1 regardless of the value of x. Therefore

n 2 3 4
∑x = 1 +x +x +x +x +… (8.6.3)

n=0

This is a geometric series in x.


2. This series is centered at c = −1 . Note how this series starts with n = 1 . We could rewrite this series starting at n = 0 with
the understanding that a = 0 , and hence the first term is 0.
0

∞ n 2 3 4 5
(x + 1) (x + 1) (x + 1) (x + 1) (x + 1)
n+1
∑(−1 ) = (x + 1) − + − + … (8.6.4)
n 2 3 4 5
n=1

3. This series is centered at c = π . Recall that 0! = 1 .


∞ 2n 2 4 6 8
(x − π) (x − π) (x − π) (x − π) (x − π)
n+1
∑(−1 ) = −1 + − + − … (8.6.5)
(2n)! 2 24 6! 8!
n=0

We introduced power series as a type of function, where a value of x is given and the sum of a series is returned. Of course, not

every series converges. For instance, in part 1 of Example 8.6.1, we recognized the series ∑ x
n
as a geometric series in x .
n=0

Theorem 60 states that this series converges only when |x| < 1.
This raises the question: "For what values of x will a given power series converge?,'' which leads us to a theorem and definition.

theorem 73: convergence of power series


Let a power series ∑ an (x − c )


n
be given. Then one of the following is true:
n=0

1. The series converges only at x = c .

8.6.1 https://math.libretexts.org/@go/page/4204
2. There is an R > 0 such that the series converges for all x in (c − R, c + R) and diverges for all x < c − R and
x > c +R .

3. The series converges for all x.

The value of R is important when understanding a power series, hence it is given a name in the following definition. Also, note that
part 2 of Theorem 73 makes a statement about the interval (c − R, c + R) , but the not the endpoints of that interval. A series
may/may not converge at these endpoints.

Definition 37: Radius and Interval of Convergence


1. The number R given in Theorem 73 is the radius of convergence of a given series. When a series converges for only
x = c , we say the radius of convergence is 0, i.e., R = 0 . When a series converges for all x, we say the series has an

infinite radius of convergence, i.e., R = ∞ .


2. The interval of convergence is the set of all values of x for which the series converges.

To find the values of x for which a given series converges, we will use the convergence tests we studied previously (especially the
Ratio Test). However, the tests all required that the terms of a series be positive. The following theorem gives us a work--around to
this problem.

theorem 74: The Radius of Convergence of a Series and Absolute Convergence


∞ ∞

The series ∑ an x
n
and ∑ ∣
n
∣an x ∣
∣ have the same radius of convergence R .
n=0 n=0

Theorem 74 allows us to find the radius of convergence R of a series by applying the Ratio Test (or any applicable test) to the
absolute value of the terms of the series. We practice this in the following example.

Example 8.6.2: Determining the radius and interval of convergence.

Find the radius and interval of convergence for each of the following series:

n

1. ∑
x

n!
n=0

n

2. ∑ (−1 )
n+1 x
n
n=1

3. n
∑ 2 (x − 3 )
n

n=0

4. ∑ n! x
n

n=0

Solution

n

1. We apply the Ratio Test to the series ∑ ∣



x

n!

∣ :
n=0

n+1 n+1

∣x /(n + 1)! ∣
∣ ∣x n! ∣
lim = lim ∣ ⋅ ∣
n n
n→∞ ∣
∣x /n! ∣

n→∞ ∣ x (n + 1)! ∣

∣ x ∣
= lim ∣ ∣
n→∞ ∣ n + 1 ∣

= 0 for all x.

2. The Ratio Test shows us that regardless of the choice of x, the series converges. Therefore the radius of convergence is
R = ∞ , and the interval of convergence is (−∞, ∞) .
∞ ∞
n n

3. We apply the Ratio Test to the series ∑ ∣


∣(−1 )
n+1 x

n
∣ = ∑ ∣
∣ ∣
x

n

∣ :
n=1 n=1

8.6.2 https://math.libretexts.org/@go/page/4204
n+1 n+1

∣x /(n + 1)∣
∣ ∣x n ∣
lim = lim ∣ ⋅ ∣
n n
n→∞ ∣
∣x /n∣

n→∞ ∣ x n+1 ∣

n
= lim |x|
n→∞ n+1

= |x|.

The Ratio Test states a series converges if the limit of |a /a | = L < 1 . We found the limit above to be |x|; therefore,
n+1 n

the power series converges when |x| < 1, or when x is in (−1, 1). Thus the radius of convergence is R = 1 .

To determine the interval of convergence, we need to check the endpoints of (−1, 1). When x = −1 , we have the opposite
of the Harmonic Series:
∞ n ∞
(−1) −1
n+1
∑(−1 ) =∑
n n
n=1 n=1

= −∞.

The series diverges when x = −1 .

∞ n
(1)
When x = 1 , we have the series ∑ (−1 )
n+1
n
, which is the Alternating Harmonic Series, which converges. Therefore
n=1

the interval of convergence is (−1, 1].


4. We apply the Ratio Test to the series ∑ ∣


n n
∣2 (x − 3 ) ∣∣ :
n=0

n+1 n+1 n+1



∣2 (x − 3 ) ∣
∣ ∣ 2n+1 (x − 3) ∣
lim = lim ∣ ⋅ ∣
n n n
n→∞ ∣ n∣ n→∞ 2 (x − 3)
∣2 (x − 3 ) ∣ ∣ ∣

= lim ∣
∣2(x − 3)∣
∣.
n→∞

According to the Ratio Test, the series converges when ∣∣2(x − 3)∣∣ < 1 ⟹ ∣∣x − 3∣∣ < 1/2 . The series is centered at 3,
and x must be within 1/2 of 3 in order for the series to converge. Therefore the radius of convergence is R = 1/2 , and we
know that the series converges absolutely for all x in (3 − 1/2, 3 + 1/2) = (2.5, 3.5) .

We check for convergence at the endpoints to find the interval of convergence. When x = 2.5, we have:
∞ ∞

n n n n
∑ 2 (2.5 − 3 ) = ∑ 2 (−1/2 )

n=0 n=0

n
= ∑(−1 ) ,

n=0

which diverges. A similar process shows that the series also diverges at x = 3.5. Therefore the interval of convergence is
(2.5, 3.5).

5. We apply the Ratio Test to ∑ ∣


n
∣n! x ∣
∣ :
n=0

n+1

∣(n + 1)! x ∣

lim = lim ∣
∣(n + 1)x ∣

n→∞ n n→∞

∣n! x ∣∣

= ∞ for all x, except x = 0.

The Ratio Test shows that the series diverges for all x except x = 0 . Therefore the radius of convergence is R = 0 .

We can use a power series to define a function:

8.6.3 https://math.libretexts.org/@go/page/4204

n
f (x) = ∑ an x (8.6.6)

n=0

where the domain of f is a subset of the interval of convergence of the power series. One can apply calculus techniques to such
functions; in particular, we can find derivatives and antiderivatives.

theorem 75: Derivatives and Indefinite Integrals of Power Series Functions


Let f (x) = ∑ an (x − c )
n
be a function defined by a power series, with radius of convergence R .
n=0

1. f (x) is continuous and differentiable on (c − R, c + R) .


2. f ′
(x) = ∑ an ⋅ n ⋅ (x − c )
n−1
, with radius of convergence R .
n=1
∞ n+1
(x−c)
3. ∫ f (x) dx = C + ∑ a n
n+1
, with radius of convergence R .
n=0

A few notes about Theorem 75:


1. The theorem states that differentiation and integration do not change the radius of convergence. It does not state anything about
the interval of convergence. They are not always the same.
2. Notice how the summation for f (x) starts with n = 1 . This is because the constant term a of f (x) goes to 0.

0

3. Differentiation and integration are simply calculated term--by--term using the Power Rules.

Example 8.6.3: Derivatives and indefinite integrals of power series


Let f (x) = ∑ x
n
. Find f ′
(x) and F (x) = ∫ f (x) dx , along with their respective intervals of convergence.
n=0

Solution
We find the derivative and indefinite integral of f (x), following Theorem 75.

1. f ′
(x) = ∑ nx
n−1
= 1 + 2x + 3 x
2
+ 4x
3
+⋯ .
n=1

In Example 8.6.1, we recognized that ∑ x


n
is a geometric series in x. We know that such a geometric series converges
n=0

when |x| < 1; that is, the interval of convergence is (−1, 1).

To determine the interval of convergence of f ′


(x) , we consider the endpoints of (−1, 1):

f (−1) = 1 − 2 + 3 − 4 + ⋯ , which diverges. (8.6.7)


f (1) = 1 + 2 + 3 + 4 + ⋯ , which diverges. (8.6.8)

Therefore, the interval of convergence of f ′


(x) is (−1, 1).

n+1 2 3

2. F (x) = ∫ f (x) dx = C + ∑ x

n+1
= C +x +
x

2
+
x

3
+⋯
n=0

To find the interval of convergence of F (x), we again consider the endpoints of (−1, 1):
F (−1) = C − 1 + 1/2 − 1/3 + 1/4 + ⋯ (8.6.9)

The value of C is irrelevant; notice that the rest of the series is an Alternating Series that whose terms converge to 0. By the
Alternating Series Test, this series converges. (In fact, we can recognize that the terms of the series after C are the opposite
of the Alternating Harmonic Series. We can thus say that F (−1) = C − ln 2 .)

8.6.4 https://math.libretexts.org/@go/page/4204
F (1) = C + 1 + 1/2 + 1/3 + 1/4 + ⋯ (8.6.10)

Notice that this summation is C + the Harmonic Series, which diverges. Since F converges for x = −1 and diverges for
x = 1 , the interval of convergence of F (x) is [−1, 1).

The previous example showed how to take the derivative and indefinite integral of a power series without motivation for why we
care about such operations. We may care for the sheer mathematical enjoyment "that we can'', which is motivation enough for
many. However, we would be remiss to not recognize that we can learn a great deal from taking derivatives and indefinite integrals.

Recall that f (x) = ∑ x


n
in Example 8.6.3 is a geometric series. According to Theorem 60, this series converges to 1/(1 − x)
n=0

when |x| < 1. Thus we can say



1
n
f (x) = ∑ x = , on (−1, 1). (8.6.11)
1 −x
n=0

Integrating the power series, (as done in Example 8.6.3,) we find


∞ n+1
x
F (x) = C1 + ∑ , (8.6.12)
n+1
n=0

while integrating the function f (x) = 1/(1 − x) gives


F (x) = − ln |1 − x| + C2 . (8.6.13)

Equating Equations 8.6.12 and 8.6.13, we have


∞ n+1
x
F (x) = C1 + ∑ = − ln |1 − x| + C2 . (8.6.14)
n+1
n=0

Letting x = 0 , we have F (0) = C 1 = C2 . This implies that we can drop the constants and conclude
∞ n+1
x
∑ = − ln |1 − x|. (8.6.15)
n+1
n=0

We established in Example 8.6.3 that the series on the left converges at x = −1 ; substituting x = −1 on both sides of the above
equality gives
1 1 1 1
−1 + − + − + ⋯ = − ln 2. (8.6.16)
2 3 4 5

On the left we have the opposite of the Alternating Harmonic Series; on the right, we have − ln 2 . We conclude that
1 1 1
1− + − + ⋯ = ln 2. (8.6.17)
2 3 4

Important: We stated in Key Idea 31 (in Section 8.2) that the Alternating Harmonic Series converges to ln 2, and referred to this
fact again in Section 8.5. However, we never gave an argument for why this was the case. The work above finally shows how we
conclude that the Alternating Harmonic Series converges to ln 2.
We use this type of analysis in the next example.

Example 8.6.4: Analyzing power series functions



n

Let f (x) = ∑
x

n!
. Find f ′
(x) and ∫ f (x) dx , and use these to analyze the behavior of f (x).
n=0

Solution
We start by making two notes: first, in Example 8.6.2, we found the interval of convergence of this power series is (−∞, ∞).
Second, we will find it useful later to have a few terms of the series written out:

8.6.5 https://math.libretexts.org/@go/page/4204
∞ n 2 3 4
x x x x
∑ = 1 +x + + + +⋯ (8.6.18)
n! 2 6 24
n=0

We now find the derivative:


∞ n−1
x

f (x) = ∑ n
n!
n=1

∞ n−1 2
x x
=∑ = 1 +x + +⋯ .
(n − 1)! 2!
n=1

Since the series starts at n = 1 and each term refers to (n − 1), we can re-index the series starting with n = 0:
∞ n
x
=∑
n!
n=0

= f (x).

We found the derivative of f (x) is f (x). The only functions for which this is true are of the form y = ce for some constant c . x

As f (0) = 1 (see Equation 8.6.18), c must be 1. Therefore we conclude that


∞ n
x x
f (x) = ∑ =e (8.6.19)
n!
n=0

for all x.
We can also find ∫ f (x) dx :
∞ n+1
x
∫ f (x)dx = C + ∑
n!(n + 1)
n=0

∞ n+1
x
= C +∑
(n + 1)!
n=0

We write out a few terms of this last series:


∞ n+1 2 3 4
x x x x
C +∑ = C +x + + + +⋯ (8.6.20)
(n + 1)! 2 6 24
n=0

The integral of f (x) differs from f (x) only by a constant, again indicating that f (x) = e . x

Example 8.6.4 and the work following Example 8.6.3 established relationships between a power series function and "regular''
functions that we have dealt with in the past. In general, given a power series function, it is difficult (if not impossible) to express
the function in terms of elementary functions. We chose examples where things worked out nicely.
In this section's last example, we show how to solve a simple differential equation with a power series.\\

Example 8.6.5: Solving a differential equation with a power series.

Give the first 4 terms of the power series solution to y ′


= 2y , where y(0) = 1 .
Solution
The differential equation y = 2y describes a function y = f (x) where the derivative of y is twice y and y(0) = 1 . This is a

rather simple differential equation; with a bit of thought one should realize that if y = C e , then y = 2C e , and hence 2x ′ 2x

y = 2y . By letting C = 1 we satisfy the initial condition of y(0) = 1 .


Let's ignore the fact that we already know the solution and find a power series function that satisfies the equation. The solution
we seek will have the form

n 2 3
f (x) = ∑ an x = a0 + a1 x + a2 x + a3 x +⋯ (8.6.21)

n=0

8.6.6 https://math.libretexts.org/@go/page/4204
for unknown coefficients a . We can find f
n

(x) using Theorem 75:

′ n−1 2 3
f (x) = ∑ an ⋅ n ⋅ x = a1 + 2 a2 x + 3 a3 x + 4 a4 x ⋯. (8.6.22)

n=1

Since f ′
, we have
(x) = 2f (x)

2 3 2 3
a1 + 2 a2 x + 3 a3 x + 4 a4 x ⋯ = 2(a0 + a1 x + a2 x + a3 x +⋯ )

2 3
= 2 a0 + 2 a1 x + 2 a2 x + 2 a3 x +⋯

The coefficients of like powers of x must be equal, so we find that


a1 = 2 a0 , 2 a2 = 2 a1 , 3 a3 = 2 a2 , 4 a4 = 2 a3 , etc. (8.6.23)

The initial condition y(0) = f (0) = 1 indicates that a 0 =1 ; with this, we can find the values of the other coefficients:
a0 = 1 and a1 = 2 a0 ⇒ a1 = 2;

a1 = 2 and 2 a2 = 2 a1 ⇒ a2 = 4/2 = 2;

a2 = 2 and 3 a3 = 2 a2 ⇒ a3 = 8/(2 ⋅ 3) = 4/3;

a3 = 4/3 and 4 a4 = 2 a3 ⇒ a4 = 16/(2 ⋅ 3 ⋅ 4) = 2/3.

Thus the first 5 terms of the power series solution to the differential equation y ′
= 2y is
4 2
2 3 4
f (x) = 1 + 2x + 2 x + x + x +⋯ (8.6.24)
3 3

In Section 8.8, as we study Taylor Series, we will learn how to recognize this series as describing y = e . 2x

Our last example illustrates that it can be difficult to recognize an elementary function by its power series expansion. It is far easier
to start with a known function, expressed in terms of elementary functions, and represent it as a power series function. One may
wonder why we would bother doing so, as the latter function probably seems more complicated. In the next two sections, we show
both how to do this and why such a process can be beneficial.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 8.6: Power Series is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et al.
via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

8.6.7 https://math.libretexts.org/@go/page/4204
8.7: Taylor Polynomials
Consider a function y = f (x) and a point (c, f (c)). The derivative, f (c), gives the instantaneous rate of change of f at x = c . Of all lines

that pass through the point (c, f (c)), the line that best approximates f at this point is the tangent line; that is, the line whose slope (rate of
change) is f (c).

In Figure 8.7.1, we see a function y = f (x) graphed. The table below the graph shows that f (0) = 2 and f (0) = 1 ; therefore, the tangent

line to f at x = 0 is p (x) = 1(x − 0) + 2 = x + 2 . The tangent line is also given in the figure. Note that "near'' x = 0 , p (x) ≈ f (x); that
1 1

is, the tangent line approximates f well.

Figure 8.7.1 : Plotting y = f (x) and a table of derivatives of f evaluated 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, p (x), that does match the concavity without much difficulty, though. The table in Figure 8.16 gives the
2

following information:
′ ′′
f (0) = 2 f (0) = 1 f (0) = 2. (8.7.1)

Therefore, we want our polynomial p 2 (x) to have these same properties. That is, we need
′ ′′
p2 (0) = 2 p (0) = 1 p (0) = 2. (8.7.2)
2 2

This is simply an initial-value problem. We can solve this using the techniques first described in Section 5.1. To keep p2 (x) as simple as
possible, we'll assume that not only p (0) = 2 , but that p (x) = 2 . That is, the second derivative of p is constant.
′′
2
′′
2 2

If ′′
p (x) = 2
2
, then p (x) = 2x + C for some constant C . Since we have determined that p (0) = 1 , we find that C = 1 and so

2

2

(x) = 2x + 1 . Finally, we can compute p (x) = x + x + C . Using our initial values, we know p (0) = 2 so C = 2. We conclude that
′ 2
p 2 2
2

(x) = x + x + 2. This function is plotted with f in Figure 8.7.2.


2
p2

Figure 8.7.2 : Plotting f (blue), p (red) and p (light red).


2 4

We can repeat this approximation process by creating polynomials of higher degree that match more of the derivatives of f at x = 0 . In
general, a polynomial of degree n can be created to match the first n derivatives of f . Figure 8.7.2 also shows
p (x) = −x /2 − x /6 + x + x + 2 , whose first four derivatives at 0 match those of f . (Using the table in Figure 8.7.1, start with
4 3 2
4

(4)
p
4
(x) = −12 and solve the related initial-value problem.)
As we use more and more derivatives, our polynomial approximation to f gets better and better. In this example, the interval on which the
approximation is "good'' gets bigger and bigger. Figure 8.7.3 shows p (x); we can visually affirm that this polynomial approximates f very
13

well on [−2, 3].

8.7.1 https://math.libretexts.org/@go/page/4673
Figure 8.7.3 : Plotting f and p . 13

The polynomial p 13 (x) is not particularly "nice''. It is


13 12 11 10 9 8 7 6 5 4 3
16901x 13x 1321x 779x 359x x 139x 11x 19x x x
p13 = + − − − + + + − − − (8.7.3)
6227020800 1209600 39916800 1814400 362880 240 5040 360 120 2 6
2
+x + x + 2.

The polynomials we have created are examples of Taylor polynomials, named after the British mathematician Brook Taylor who made
important discoveries about such functions. While we created the above Taylor polynomials by solving initial-value problems, it can be
shown that Taylor polynomials follow a general pattern that make their formation much more direct. This is described in the following
definition.

Definition 38: Taylor Polynomials & Maclaurin PolynomiaLS


Let f be a function whose first n derivatives exist at x = c .
1. The Taylor polynomial of degree n of f at x = c is
′′ ′′′ (n)
f (c) f (c) f (c)
′ 2 3 n
pn (x) = f (c) + f (c)(x − c) + (x − c ) + (x − c ) +⋯ + (x − c ) . (8.7.4)
2! 3! n!

2. A special case of the Taylor polynomial is the Maclaurin polynomial, where c = 0 . That is, the Maclaurin polynomial of degree n
of f is
′′ ′′′ (n)
f (0) f (0) f (0)
′ 2 3 n
pn (x) = f (0) + f (0)x + x + x +⋯ + x . (8.7.5)
2! 3! n!

We will practice creating Taylor and Maclaurin polynomials in the following examples.

Example 8.7.1: Finding and using Maclaurin polynomials


1. Find the n Maclaurin polynomial for f (x) = e .
th x

2. Use p (x) to approximate the value of e .


5

Solution

Figure 8.7.4 The derivatives of f (x) = e evaluated at x = 0 . x

1. We start with creating a table of the derivatives of e evaluated at x = 0 . In this particular case, this is relatively simple, as shown in
x

Figure 8.19.

By the definition of the Maclaurin series, we have


′′ ′′′ n
f (0) f (0) f (0)
′ 2 3 n
pn (x) = f (0) + f (0)x + x + x +⋯ + x
2! 3! n!
1 1 1 1
2 3 4 n
= 1 +x + x + x + x +⋯ + x .
2 6 24 n!

2. Using our answer from part 1, we have


1 2
1 3
1 4
1 5
p5 = 1 + x + x + x + x + x . (8.7.6)
2 6 24 120

8.7.2 https://math.libretexts.org/@go/page/4673
To approximate the value of e , note that e = e 1
= f (1) ≈ p5 (1). It is very straightforward to evaluate p 5 (1) :
1 1 1 1 163
p5 (1) = 1 + 1 + + + + = ≈ 2.71667. (8.7.7)
2 6 24 120 60

A plot of f (x) = e and p x


5 (x) is given in Figure 8.7.5.

Figure 8.7.5 : A plot of f (x) = e and its 5th degree Maclaurin polynomial p
x
5 (x) .

Example 8.7.2: Finding and using Taylor polynomials


1. Find the n Taylor polynomial of y = ln x at x = 1 .
th

2. Use p (x) to approximate the value of ln 1.5.


6

3. Use p (x) to approximate the value of ln 2.


6

Solution

Figure 8.7.6 : Derivatives of ln x evaluated at x = 1 .


1. We begin by creating a table of derivatives of ln x evaluated at x = 1 . While this is not as straightforward as it was in the previous
example, a pattern does emerge, as shown in Figure 8.7.6.
Using Definition 38, we have
′′ ′′′ n
f (c) f (c) f (c)
′ 2 3 n
pn (x) = f (c) + f (c)(x − c) + (x − c ) + (x − c ) +⋯ + (x − c )
2! 3! n!
n+1
1 1 1 (−1)
2 3 4 n
= 0 + (x − 1) − (x − 1 ) + (x − 1 ) − (x − 1 ) +⋯ + (x − 1 ) .
2 3 4 n

Note how the coefficients of the (x − 1) terms turn out to be "nice.''


2. We can compute p (x) using our work above:
6

1 1 1 1 1
2 3 4 5 6
p6 (x) = (x − 1) − (x − 1 ) + (x − 1 ) − (x − 1 ) + (x − 1 ) − (x − 1 ) . (8.7.8)
2 3 4 5 6

Since p 6 (x) approximates ln x well near x = 1 , we approximate ln 1.5 ≈ p 6 (1.5) :


1 1 1
2 3 4
p6 (1.5) = (1.5 − 1) − (1.5 − 1 ) + (1.5 − 1 ) − (1.5 − 1 ) +⋯
2 3 4

1 5
1 6
⋯+ (1.5 − 1 ) − (1.5 − 1 )
5 6
259
=
640

≈ 0.404688.

This is a good approximation as a calculator shows that ln 1.5 ≈ 0.4055. Figure 8.7.7 plots y = ln x with y = p 6 (x) . We can see
that ln 1.5 ≈ p (1.5).6

8.7.3 https://math.libretexts.org/@go/page/4673
Figure 8.7.7 : A plot of y = ln x and its 6th degree Taylor polynomial at x = 1 .
3. We approximate ln 2 with p 6 (2):
1 2
1 3
1 4
p6 (2) = (2 − 1) − (2 − 1 ) + (2 − 1 ) − (2 − 1 ) +⋯
2 3 4
1 1
5 6
⋯+ (2 − 1 ) − (2 − 1 )
5 6
1 1 1 1 1
=1− + − + −
2 3 4 5 6
37
=
60

≈ 0.616667.

This approximation is not terribly impressive: a hand held calculator shows that ln 2 ≈ 0.693147. The graph in Figure 8.22 shows
that p (x) provides less accurate approximations of ln x as x gets close to 0 or 2.
6

Surprisingly enough, even the 20 th


degree Taylor polynomial fails to approximate ln x for x > 2 , as shown in Figure 8.7.8. We'll
soon discuss why this is.

Figure 8.7.8 : A plot of y = ln x and its 20th degree Taylor polynomial at x = 1 .

Taylor polynomials are used to approximate functions f (x) in mainly two situations:
1. When f (x) is known, but perhaps "hard'' to compute directly. For instance, we can define y = cos x as either the ratio of sides of a right
triangle ("adjacent over hypotenuse'') or with 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 operations usually
hard-wired into a computer (+, −, × and ÷).
2. When f (x) is not known, but information about its derivatives is known. This occurs more often than one might think, especially in the
study of differential equations.

Even though Taylor polynomials could be used in calculators and computers to calculate values of trigonometric functions, in practice
they generally aren't. Other more efficient and accurate methods have been developed, such as the CORDIC algorithm.

In both situations, a critical piece of information to have is "How good is my approximation?'' If we use a Taylor polynomial to compute
cos 2, how do we know how accurate the approximation is?

We had the same problem when studying Numerical Integration. Theorem 43 provided bounds on the error when using, say, Simpson's Rule
to approximate a definite integral. These bounds allowed us to determine that, for instance, using 10 subintervals provided an approximation
within ±.01 of the exact value. The following theorem gives similar bounds for Taylor (and hence Maclaurin) polynomials.

8.7.4 https://math.libretexts.org/@go/page/4673
THEOREM 76: TAYLOR'S THEOREM
1. Let f be a function whose n + 1 th
derivative exists on an interval I and let c be in I . Then, for each x in I , there exists z between x

x and c such that

′ (n)
f (c) f (c)
′ 2 n
f (x) = f (c) + f (c)(x − c) + (x − c ) +⋯ + (x − c ) + Rn (x), (8.7.9)
2! n!

(n+1)
f (zx )
where R n (x) = (x − c )
(n+1)
.
(n + 1)!
(n+1)
max ∣
∣f (z)∣

2. ∣∣R ∣ ≤
n (x)∣

∣(x − c )
(n+1)


(n + 1)!

The first part of Taylor's Theorem states that f (x) = p (x) + R (x) , where p (x) is the n order Taylor polynomial and R (x) is the
n n n
th
n

remainder, or error, in the Taylor approximation. The second part gives bounds on how big that error can be. If the (n + 1) derivative is th

large, 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 8.7.2.

Example 8.7.3: Finding error bounds of a Taylor polynomial

Use Theorem 76 to find error bounds when approximating ln 1.5 and ln 2 with p 6 (x) , the Taylor polynomial of degree 6 of f (x) = ln x
at x = 1 , as calculated in Example 8.7.2.
Solution
1. We start with the approximation of ln 1.5 with p (1.5). The theorem references an open interval I that contains both x and c . The
6

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 ∣ . In our situation, this is asking "How big can the 7
(z)∣
(n+1)
derivative of y = ln x be on the th

interval (0.9, 1.6)?'' The seventh derivative is y = −6!/x . The largest value it attains on I is about 1506. Thus we can bound the
7

error as:
(7)
max ∣
∣f (z)∣

7

∣R6 (1.5)∣
∣ ≤ ∣
∣(1.5 − 1 ) ∣

7!
1506 1
≤ ⋅
7
5040 2

≈ 0.0023.

We computed p (1.5) = 0.404688; using a calculator, we find ln 1.5 ≈ 0.405465, so the actual error is about 0.000778, which is
6

less than our bound of 0.0023. This affirms Taylor's Theorem; the theorem states that our approximation would be within about 2
thousandths of the actual value, whereas the approximation was actually closer.
2. We again find an interval I that contains both c = 1 and x = 2 ; we choose I = (0.9, 2.1). The maximum value of the seventh
derivative of f on this interval is again about 1506 (as the largest values come near x = 0.9). Thus
(7)
max ∣
∣f (z)∣

7

∣R6 (2)∣
∣ ≤ ∣
∣(2 − 1 ) ∣

7!
1506
7
≤ ⋅1
5040

≈ 0.30.

This bound is not as nearly as good as before. Using the degree 6 Taylor polynomial at x = 1 will bring us within 0.3 of the correct
answer. As p (2) ≈ 0.61667, our error estimate guarantees that the actual value of ln 2 is somewhere between 0.31667 and
6

0.91667. These bounds are not particularly useful.

8.7.5 https://math.libretexts.org/@go/page/4673
In reality, our approximation was only off by about 0.07. However, we are approximating ostensibly because we do not know the real
answer. In order to be assured that we have a good approximation, we would have to resort to using a polynomial of higher degree.

We practice again. This time, we use Taylor's theorem to find n that guarantees our approximation is within a certain amount.

Example 8.7.4: Finding sufficiently accurate Taylor polynomials

Find n such that the n th


Taylor polynomial of f (x) = cos x at x = 0 approximates cos 2 to within 0.001 of the actual answer. What is
p (2)?
n

Solution
Following Taylor's theorem, we need bounds on the size of the derivatives of f (x) = cos x. In the case of this trigonometric function,
this is easy. All derivatives of cosine are ± sin x or ± cos x. In all cases, these functions are never greater than 1 in absolute value. We
want the error to be less than 0.001. To find the appropriate n , consider the following inequalities:
(n+1)
max ∣
∣f (z)∣

(n+1)

∣(2 − 0 ) ∣
∣ ≤ 0.001
(n + 1)!

1 (n+1)
⋅2 ≤ 0.001
(n + 1)!

8+1
2
We find an n that satisfies this last inequality with trial-and-error. When n =8 , we have ≈ 0.0014 ; when n =9 , we have
(8 + 1)!
9+1
2
≈ 0.000282 < 0.001 . Thus we want to approximate cos 2 with p 9 (2) .\\
(9 + 1)!

We now set out to compute p . We again need a table of the derivatives of f (x) = cos x evaluated at x = 0 . A table of these values
9 (x)

is given in Figure 8.7.8.

Figure 8.7.9 : A table of the derivatives of f (x) = cosx evaluated at x = 0 .


Notice how the derivatives, evaluated at x = 0 , follow a certain pattern. All the odd powers of x in the Taylor polynomial will disappear
as their coefficient is 0. While our error bounds state that we need p (x), our work shows that this will be the same as p (x).
9 8

Since we are forming our polynomial at x = 0 , we are creating a Maclaurin polynomial, and:
′′ ′′′ (8)
f (0) f (0) f
′ 2 3 8
p8 (x) = f (0) + f (0)x + x + x +⋯ + x
2! 3! 8!
1 2
1 4
1 6
1 8
=1− x + x − x + x
2! 4! 6! 8!

We finally approximate cos 2:


131
cos 2 ≈ p8 (2) = − ≈ −0.41587.
315

Our error bound guarantee that this approximation is within 0.001 of the correct answer. Technology shows us that our approximation is
actually within about 0.0003 of the correct answer.
Figure 8.7.10 shows a graph of y = p 8 (x) and y = cos x . Note how well the two functions agree on about (−π, π).

8.7.6 https://math.libretexts.org/@go/page/4673
Figure 8.7.10 : A graph of f (x) = cosx and its degree 8 Maclaurin polynomial.

Example 8.7.5: Finding and using Taylor polynomials


1. Find the degree 4 Taylor polynomial, p (x), for f (x) = √−
4 x at x = 4.

2. Use p (x) to approximate √3.
4

3. Find bounds on the error when approximating √3 with p (3). 4

Solution

Figure 8.7.11 : A table of the derivatives of f (x) = √−


x evaluated at x = 4 .

1. We begin by evaluating the derivatives of f at x = 4 . This is done in Figure 8.7.11. These values allow us to form the Taylor
polynomial p (x):
4

1 −1/32 3/256 −15/2048


2 3 4
p4 (x) = 2 + (x − 4) + (x − 4 ) + (x − 4 ) + (x − 4 ) . (8.7.10)
4 2! 3! 4!


2. As p (x) ≈ √−
4 x near x = 4 , we approximate √3 with p (3) = 1.73212.
4

3. To find a bound on the error, we need an open interval that contains x = 3 and x = 4 . We set I . The largest value the
= (2.9, 4.1)

fifth derivative of f (x) = √−x takes on this interval is near x = 2.9 , at about 0.0273. Thus

0.0273 5

∣R4 (3)∣
∣ ≤ ∣
∣(3 − 4 ) ∣
∣ ≈ 0.00023. (8.7.11)
5!

This shows our approximation is accurate to at least the first 2 places after the decimal. (It turns out that our approximation is
actually accurate to 4 places after the decimal.) A graph of f (x) = √− x and p (x) is given in Figure 8.7.12. Note how the two
4

functions are nearly indistinguishable on (2, 7).

Figure 8.7.12 : A graph of f (x) = √−


x and its degree 4 Taylor polynomial at x = 4 .

Our final example gives a brief introduction to using Taylor polynomials to solve differential equations.

8.7.7 https://math.libretexts.org/@go/page/4673
Example 8.7.6: Approximating an unknown function

A function y = f (x) is unknown save for the following two facts.


1. y(0) = f (0) = 1 , and
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 p 3 (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:′

′ 2 ′ 2
y =y ⇒ y (0) = y (0).

Since y(0) = 1 , we conclude that y ′


(0) = 1 .

Now we find information about y


′′
. Starting with y

=y
2
, take derivatives of both sides, with respect to x. That means we must use
implicit differentiation.
′ 2
y =y

d d
′ 2
(y ) = (y )
dx dx
′′ ′
y = 2y ⋅ y .

Now evaluate both sides at x = 0:


′′ ′
y (0) = 2y(0) ⋅ y (0)
′′
y (0) = 2

We repeat this once more to find y ′′′


. We again use implicit differentiation; this time the Product Rule is also required.
(0)

d d
′′ ′
(y ) = (2y y )
dx dx
′′′ ′ ′ ′′
y = 2 y ⋅ y + 2y ⋅ y .

Now evaluate both sides at x = 0:


′′′ ′ 2 ′′
y (0) = 2 y (0 ) + 2y(0)y (0)
′′′
y (0) = 2 + 4 = 6

In summary, we have:
′ ′′ ′′′
y(0) = 1 y (0) = 1 y (0) = 2 y (0) = 6.

We can now form p :


3 (x)

2 6
2 3
p3 (x) = 1 + x + x + x
2! 3!
2 3
= 1 +x +x +x .

Figure 8.7.13 : A graph of y = −1/(x − 1) and y = p a (x) from Example 8.7.6.


It turns out that the differential equation we started with, y

=y
2
, where y(0) = 1 , can be solved without too much difficulty:
1
y = . Figure 8.28 shows this function plotted with p 3 (x) . Note how similar they are near x = 0 .
1 −x

8.7.8 https://math.libretexts.org/@go/page/4673
It is beyond the scope of this text to pursue error analysis when using Taylor polynomials to approximate solutions to differential equations.
This topic is often broached in introductory Differential Equations courses and usually covered 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.
Most of this chapter has been devoted to the study of infinite series. This section has taken a step back from this study, focusing instead on
finite summation of terms. In the next section, we explore Taylor Series, where we represent a function with an infinite series.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of VMI and
Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution - Noncommercial (BY-
NC) License. http://www.apexcalculus.com/

This page titled 8.7: Taylor Polynomials is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et al. via
source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

8.7.9 https://math.libretexts.org/@go/page/4673
8.8: Taylor Series
In Section 8.6, we showed how certain functions can be represented by a power series function. In 8.7, we showed how we can
approximate functions with polynomials, given that enough derivative information is available. In this section we combine these
concepts: if a function f (x) is infinitely differentiable, we show how to represent it with a power series function.

Definition 39 taylor and maclaurin series


Let f (x) have derivatives of all orders at x = c .
1. The Taylor Series of f (x), centered at c is
∞ (n)
f (c)
n
∑ (x − c ) . (8.8.1)
n!
n=0

2. Setting c = 0 gives the Maclaurin Series of f (x):


∞ (n)
f (0)
n
∑ x . (8.8.2)
n!
n=0

The difference between a Taylor polynomial and a Taylor series is the former is a polynomial, containing only a finite number of
terms, whereas the latter is a series, a summation of an infinite set of terms. When creating the Taylor polynomial of degree n for a
function f (x) at x = c ,we needed to evaluate f ,and the first n derivatives of f ,at x = c .When creating the Taylor series of f , it
helps to find a pattern that describes the n derivative of f at x = c .We demonstrate this in the next two examples.
th

Example 8.8.1: The Maclaurin series of f (x) = cos x

Find the Maclaurin series of f (x) = cos x.


Solution
In Example 8.7.4 we found the 8
th
degree Maclaurin polynomial of cos x .In doing so, we created the table shown in Figure
8.29.

Figure 8.29: A table of the derivatives of f (x) = cos x evaluated at x = 0 .


Notice how f (0) = 0 when n is odd, f (0) = 1 when
(n) (n)
n is divisible by 4,and f
(n)
(0) = −1 when n is even but not
divisible by 4. Thus the Maclaurin series of cos x is
2 4 6 8
x x x x
1− + − + −⋯ (8.8.3)
2 4! 6! 8!

We can go further and write this as a summation. Since we only need the terms where the power of x is even, we write the
power series in terms of x :
2n

∞ 2n
x
n
∑(−1 ) . (8.8.4)
(2n)!
n=0

8.8.1 https://math.libretexts.org/@go/page/4674
Example 8.8.2: The Taylor series of f (x) = ln x at x = 1

Find the Taylor series of f (x) = ln x centered at x = 1 .


Solution
Figure 8.30 shows the n th
derivative of ln x evaluated at x = 1 for n = 0, … , 5,along with an expression for the n th
term:
(n) n+1
f (1) = (−1 ) (n − 1)! for n ≥ 1. (8.8.5)

Remember that this is what distinguishes Taylor series from Taylor polynomials; we are very interested in finding a pattern for
the n term, not just finding a finite set of coefficients for a polynomial.
th

Figure 8.30: Derivatives of ln x evaluated at x = 1 .


Since f (1) = ln 1 = 0 ,we skip the first term and start the summation with n = 1 ,giving the Taylor series for ln x,centered at
x = 1 ,as

∞ ∞ n
1 (x − 1)
n+1 n n+1
∑(−1 ) (n − 1)! (x − 1 ) = ∑(−1 ) . (8.8.6)
n! n
n=1 n=1

It is important to note that Definition 39 defines a Taylor series given a function f (x); however, we cannot yet state that f (x) is
equal to its Taylor series. We will find that "most of the time'' they are equal, but we need to consider the conditions that allow us to
conclude this.
Theorem 76 states that the error between a function f (x) and its n --degree Taylor polynomial p
th
n (x) is R ,where
n (x)

(n+1)
max ∣
∣f (z)∣

(n+1)

∣Rn (x)∣
∣ ≤ ∣
∣(x − c ) ∣
∣. (8.8.7)
(n + 1)!

If R (x) goes to 0 for each


n x in an interval I as n approaches infinity, we conclude that the function is equal to its Taylor series
expansion.

theorem 77 function and taylor series equality


Let f (x) have derivatives of all orders at x =c ,let Rn (x) be as stated in Theorem 76, and let I be an interval on which the
Taylor series of f (x) converges.
If lim Rn (x) = 0 for all x in I ,then
n→∞

∞ (n)
f (c)
n
f (x) = ∑ (x − c ) on I . (8.8.8)
n!
n=0

We demonstrate the use of this theorem in an example.

Example 8.8.3: Establishing equality of a function and its Taylor series

Show that f (x) = cos x is equal to its Maclaurin series, as found in Example 8.8.1, for all x.

Solution
Given a value x,the magnitude of the error term R n (x) is bounded by

8.8.2 https://math.libretexts.org/@go/page/4674
(n+1)
max ∣
∣f (z)∣

n+1

∣Rn (x)∣
∣ ≤ ∣
∣x ∣
∣. (8.8.9)
(n + 1)!

Since all derivatives of cos x are ± sin x or ± cos x,whose magnitudes are bounded by 1,we can state
1
n+1

∣Rn (x)∣
∣ ≤ ∣
∣x ∣
∣ (8.8.10)
(n + 1)!

which implies
n+1 n+1
|x | |x |
− ≤ Rn (x) ≤ . (8.8.11)
(n + 1)! (n + 1)!

n+1

For any x, lim x


=0 . Applying the Squeeze Theorem to Equation 8.8.11 , we conclude that lim Rn (x) = 0 for all
n→∞ (n+1)! n→∞

x ,and hence
∞ 2n
x
n
cos x = ∑(−1 ) for all x. (8.8.12)
(2n)!
n=0

It is natural to assume that a function is equal to its Taylor series on the series' interval of convergence, but this is not the case. In
order to properly establish equality, one must use Theorem 77. This is a bit disappointing, as we developed beautiful techniques for
determining the interval of convergence of a power series, and proving that R (x) → 0 can be cumbersome as it deals with high
n

order derivatives of the function.


There is good news. A function f (x) that is equal to its Taylor series, centered at any point the domain of f (x),is said to be an
analytic function, and most, if not all, functions that we encounter within this course are analytic functions. Generally speaking,
any function that one creates with elementary functions (polynomials, exponentials, trigonometric functions, etc.) that is not
piecewise defined is probably analytic. For most functions, we assume the function is equal to its Taylor series on the series'
interval of convergence and only use Theorem 77 when we suspect something may not work as expected.
We develop the Taylor series for one more important function, then give a table of the Taylor series for a number of common
functions.

Example 8.8.4: The Binomial Series

Find the Maclaurin series of f (x) = (1 + x) ,k ≠ 0 .


k

Solution
When k is a positive integer, the Maclaurin series is finite. For instance, when k = 4 ,we have
4 2 3 4
f (x) = (1 + x ) = 1 + 4x + 6 x + 4x +x . (8.8.13)

The coefficients of x when k is a positive integer are known as the binomial coefficients, giving the series we are developing
its name.
−−−−−
When k = 1/2 ,we have f (x) = √1 + x .Knowing a series representation of this function would give a useful way of

−−
approximating √1.3,for instance.
To develop the Maclaurin series for f (x) = (1 + x)
k
for any value of k ≠0 ,we consider the derivatives of f evaluated at
x = 0:

8.8.3 https://math.libretexts.org/@go/page/4674
Thus the Maclaurin series for f (x) = (1 + x) is k

k(k − 1) k(k − 1)(k − 2) k(k − 1) ⋯ (k − (n − 1))


1 +k+ + +… + +… (8.8.14)
2! 3! n!

It is important to determine the interval of convergence of this series. With

k(k − 1) ⋯ (k − (n − 1))
n
an = x , (8.8.15)
n!

we apply the Ratio Test:

| an+1 | ∣ k(k − 1) ⋯ (k − n) ∣ ∣ k(k − 1) ⋯ (k − (n − 1)) ∣


n+1 n
lim = lim ∣ x ∣/∣ x ∣
n→∞ | an | n→∞ ∣ (n + 1)! ∣ ∣ n! ∣

∣ k−n ∣
= lim ∣ x∣
n→∞ ∣ n ∣

= |x|.

The series converges absolutely when the limit of the Ratio Test is less than 1; therefore, we have absolute convergence when
|x| < 1.

While outside the scope of this text, the interval of convergence depends on the value of k .When k > 0 ,the interval of
convergence is [−1, 1].When −1 < k < 0 ,the interval of convergence is [−1, 1).If k ≤ −1 ,the interval of convergence is
(−1, 1).

We learned that Taylor polynomials offer a way of approximating a "difficult to compute'' function with a polynomial. Taylor series
offer a way of exactly representing a function with a series. One probably can see the use of a good approximation; is there any use
of representing a function exactly as a series?
While we should not overlook the mathematical beauty of Taylor series (which is reason enough to study them), there are practical
uses as well. They provide a valuable tool for solving a variety of problems, including problems relating to integration and
differential equations.
In Key Idea 32 (on the following page) we give a table of the Taylor series of a number of common functions. We then give a
theorem about the "algebra of power series,'' that is, how we can combine power series to create power series of new functions.
This allows us to find the Taylor series of functions like f (x) = e cos x by knowing the Taylor series of e and cos x.
x x

Before we investigate combining functions, consider the Taylor series for the arctangent function (see Key Idea 32). Knowing that
(1) = π/4 ,we can use this series to approximate the value of π:
−1
tan

π −1
1 1 1 1
= tan (1) = 1 − + − + −⋯ (8.8.16)
4 3 5 7 9

1 1 1 1
π = 4 (1 − + − + − ⋯) (8.8.17)
3 5 7 9

Unfortunately, this particular expansion of π converges very slowly. The first 100 terms approximate π as ,which is not
3.13159

particularly good.

KEY IDEA 32 IMPORTANT TAYLOR SERIES EXPASIONS

8.8.4 https://math.libretexts.org/@go/page/4674
THEOREM 78 ALGEBRA OF POWER SERIES
Let f (x) = ∑ ∞

n=0
an x
n
and g(x) = ∑ ∞

n=0
bn x
n
converge absolutely for |x| < R ,and let h(x) be continuous.

1. f (x) ± g(x) = ∑ (a ± b )x \quad for |x| < R .
n=0 n n
n

∞ ∞ ∞
2. f (x)g(x) = (∑ a x ) (∑
n=0
b x ) =∑
n
n
(a b n=0 n
n
n=0 0 n + a1 bn−1 + … an b0 )x
n
for |x| < R .
∞ n
3. f (h(x)) = ∑ a (h(x))
n=0
for |h(x)| < R .
n

Example 8.8.5: Combining Taylor series

Write out the first 3 terms of the Taylor Series for f (x) = e x
cos x using Key Idea 32 and Theorem 78.
Solution
Key Idea 32 informs us that
2 3 2 4
x
x x x x
e = 1 +x + + +⋯ and cos x = 1 − + +⋯ . (8.8.18)
2! 3! 2! 4!

Applying Theorem 78, we find that


2 3 2 4
x
x x x x
e cos x = (1 + x + + + ⋯) (1 − + + ⋯) . (8.8.19)
2! 3! 2! 4!

Distribute the right hand expression across the left: (8.8.20)

2 4 2 4 2 2 4
x x x x x x x
= 1 (1 − + + ⋯) + x (1 − + + ⋯) + (1 − + + ⋯) (8.8.21)
2! 4! 2! 4! 2! 2! 4!

3 2 4 4 2 4
x x x x x x
+ (1 − + + ⋯) + (1 − + + ⋯) + ⋯ (8.8.22)
3! 2! 4! 4! 2! 4!

Distribute again and collect like terms. (8.8.23)

3 4 5 7
x x x x
= 1 +x − − − + +⋯ (8.8.24)
3 6 30 630

While this process is a bit tedious, it is much faster than evaluating all the necessary derivatives of e x
cos x and computing the
Taylor series directly.

8.8.5 https://math.libretexts.org/@go/page/4674
Because the series for e and cos x both converge on (−∞, ∞),so does the series expansion for e
x x
.
cos x

Example 8.8.6: Creating new Taylor series

Use Theorem 78 to create series for y = sin(x 2


) and y = ln(√−
x) .

Solution
Given that
∞ 2n+1 3 5 7
n
x x x x
sin x = ∑(−1 ) =x− + − +⋯ , (8.8.25)
(2n + 1)! 3! 5! 7!
n=0

we simply substitute x for x in the series, giving


2

∞ 2 2n+1 6 10 14
(x ) x x x
2 n 2
sin(x ) = ∑(−1 ) =x − + − ⋯. (8.8.26)
(2n + 1)! 3! 5! 7!
n=0

Since the Taylor series for sin x has an infinite radius of convergence, so does the Taylor series for sin(x ). 2


The Taylor expansion for ln x given in Key Idea 32 is centered at x =1 ,so we will center the series for ln(√x ) at x =1 as
well.
With
∞ n 2 3
(x − 1) (x − 1) (x − 1)
n+1
ln x = ∑(−1 ) = (x − 1) − + −⋯ , (8.8.27)
n 2 3
n=1


we substitute √x for x to obtain
∞ − n − 2 − 3
(√x − 1 ) (√x − 1 ) (√x − 1 )
− n+1 −
ln(√x ) = ∑(−1 ) = (√x − 1) − + −⋯ . (8.8.28)
n 2 3
n=1

While this is not strictly a power series, it is a series that allows us to study the function ln(√− x ).Since the interval of

convergence of ln x is (0, 2],and the range of √x on (0, 4] is (0, 2],the interval of convergence of this series expansion of

ln(√x ) is (0, 4].

Note: In Example 8.8.6, one could create a series for ln(√− −


x ) by simply recognizing that ln(√x ) = ln(x ) = 1/2 ln x ,and
1/2

hence multiplying the Taylor series for ln x by 1/2.This example was chosen to demonstrate other aspects of series, such as the
fact that the interval of convergence changes.

Example 8.8.7: Using Taylor series to evaluate definite integrals


2 1 2

Use the Taylor series of e −x


to evaluate ∫ 0
e
−x
dx .

Solution
We learned, when studying Numerical Integration, that e does not have an antiderivative expressible in terms of elementary
2
−x

functions. This means any definite integral of this function must have its value approximated, and not computed exactly.
2

We can quickly write out the Taylor series for e −x


using the Taylor series of e : x

8.8.6 https://math.libretexts.org/@go/page/4674
∞ n 2 3
x
x x x
e =∑ = 1 +x + + +⋯
n! 2! 3!
n=0

and so
∞ 2 n
2 (−x )
−x
e =∑
n!
n=0

∞ 2n
x
n
= ∑(−1 )
n!
n=0

4 6
2
x x
= 1 −x + − +⋯ .
2! 3!

We use Theorem 75 to integrate:


3 5 7 2n+1
2
−x
x x x n
x
∫ e dx = C + x − + − + ⋯ + (−1 ) +⋯ (8.8.29)
3 5 ⋅ 2! 7 ⋅ 3! (2n + 1)n!

This is the antiderivative of e ;while we can write it out as a series, we cannot write it out in terms of elementary functions.
2
−x

1
We can evaluate the definite integral ∫ e dx using this antiderivative; substituting 1 and 0 for x and subtracting gives
2
−x

1
2 1 1 1 1
−x
∫ e dx = 1 − + − + ⋯. (8.8.30)
0
3 5 ⋅ 2! 7 ⋅ 3! 9 ⋅ 4!

Summing the 5 terms shown above give the approximation of 0.74749. Since this is an alternating series, we can use the
Alternating Series Approximation Theorem, (Theorem 71), to determine how accurate this approximation is. The next term of
the series is 1/(11 ⋅ 5!) ≈ 0.00075758.Thus we know our approximation is within 0.00075758 of the actual value of the
integral. This is arguably much less work than using Simpson's Rule to approximate the value of the integral.

Example 8.8.8: Using Taylor series to solve differential equations

Solve the differential equation y = 2y in terms of a power series, and use the theory of Taylor series to recognize the solution

in terms of an elementary function.


Solution
We found the first 5 terms of the power series solution to this differential equation in Example 8.6.5 in Section 8.6. These are:
4 8 4 16 2
a0 = 1, a1 = 2, a2 = = 2, a3 = = , a4 = = . (8.8.31)
2 2⋅3 3 2⋅3⋅4 3

We include the "unsimplified'' expressions for the coefficients found in Example 8.6.5 as we are looking for a pattern. It can be
shown that a = 2 /n! .Thus the solution, written as a power series, is
n
n

∞ n ∞ n
2 (2x)
n
y =∑ x =∑ . (8.8.32)
n! n!
n=0 n=0

Using Key Idea 32 and Theorem 78, we recognize f (x) = e : 2x

∞ n ∞ n
x (2x)
x 2x
e =∑ ⇒ e =∑ . (8.8.33)
n! n!
n=0 n=0

Finding a pattern in the coefficients that match the series expansion of a known function, such as those shown in Key Idea 32, can
be difficult. What if the coefficients in the previous example were given in their reduced form; how could we still recover the
function y = e ?
2x

Suppose that all we know is that


4 2
a0 = 1, a1 = 2, a2 = 2, a3 = , a4 = . (8.8.34)
3 3

Definition 39 states that each term of the Taylor expansion of a function includes an n!.This allows us to say that

8.8.7 https://math.libretexts.org/@go/page/4674
b2 4 b3 2 b4
a2 = 2 = , a3 = = , and a4 = = (8.8.35)
2! 3 3! 3 4!

for some values b ,b and b .


2 3 4

Solving for these values, we see that ,


b2 = 4 b3 = 8 and b4 = 16 .That is, we are recovering the pattern we had previously seen,
allowing us to write
∞ ∞
n
bn n
f (x) = ∑ an x =∑ x
n!
n=0 n=0

4 2
8 3
16 4
= 1 + 2x + x + x + x +⋯
2! 3! 4!

From here it is easier to recognize that the series is describing an exponential function.
There are simpler, more direct ways of solving the differential equation y = 2y .We applied power series techniques to this

equation to demonstrate its utility, and went on to show how sometimes we are able to recover the solution in terms of elementary
functions using the theory of Taylor series. Most differential equations faced in real scientific and engineering situations are much
more complicated than this one, but power series can offer a valuable tool in finding, or at least approximating, the solution.
This chapter introduced sequences, which are ordered lists of numbers, followed by series, wherein we add up the terms of a
sequence. We quickly saw that such sums do not always add up to "infinity,'' but rather converge. We studied tests for convergence,
then ended the chapter with a formal way of defining functions based on series. Such "series--defined functions'' are a valuable tool
in solving a number of different problems throughout science and engineering.
Coming in the next chapters are new ways of defining curves in the plane apart from using functions of the form y = f (x).Curves
created by these new methods can be beautiful, useful, and important.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 8.8: Taylor Series is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et al.
via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

8.8.8 https://math.libretexts.org/@go/page/4674
8.E: Applications of Sequences and Series (Exercises)
8.1: Sequences
Terms and Concepts
1. Use your own words to define a sequence.
2. The domain of a sequence is the _____ numbers.
3. Use your own words to describe the range of a sequence.
4. Describe what it means for a sequence to be bounded.

Problems
In Exercises 5-8, give the first five terms of the given sequence.
n

5. a n ={
(n+1)!
4
}

n
6. bn = {(−
3

2
) }

n+1

7. c n = {−
n

n+2
}

n 2
1+√5 1−√5
8. d n ={
√5
1
((
2
) −(
2
) )}

In Exercises 9-12, determine the n term of the given sequence. th

9. 4, 7, 10, 13, 16, ...


10. 3, −
3

2
,
3

4
, −
3

8
,...

11. 10, 20, 40, 80, 160, ...


12. 1, 1,
1

2
,
1

6
,
24
1
,
1

120
,...

In Exercises 13-16, use the following information to determine the limit of the given sequences.
n
2 −20
an = { n
}; lim an = 1
2 n→∞

n
2 2
bn = {(1 + ) }; lim bn = e
n
n→∞
n
2 −20
cn = { n
}; lim cn = 0
2 n→∞

13. a n ={
2 −20

7⋅2
n
}

14. a n = 3 bn − an

n
15. a n = {∑(3/n)(1 =
2

n
) }

2n
16. a n = {(1 =
2

n
) }

In Exercises 17-28, determine whether the sequence converges or diverges. If convergent, give the limit of the sequence.

17. a n = {(−1 )
n

n+1
n
}

18. a n ={
4 n −n+5

3 n +1
2
}

19. a n ={
4

5
n
}

n−1
20. a n ={
n

n

n−1
}, n ≥2

21. a n = {ln(n)}

8.E.1 https://math.libretexts.org/@go/page/9979
22. an ={
3n
}
√n2 +1

n
23. an = {(1 +
1

n
) }

24. an = {5 −
n
1
}

n+1
(−1)
25. an ={
n
}

26. an ={
1.1

n
}

27. an ={
n+1
2n
}

28. an = {(−1 )
n
n
2 −1
n
}

In Exercises 29-34, determine whether the sequence is bounded, bounded above, bounded below, or none of the above.
29. an = ∑n

30. an = tan n

31. an = (−1 )
n 3n−1
n

32. an ={
3 n −1

n
}

33. an = n cos n

34. an =2
n
− n!

In Exercises 35-38, determine whether the sequence is monotonically increasing or decreasing. If it is not, determine if there
is an m such that it is monotonic for all n ≥ m .

35. an ={
n+2
n
}

36. an ={
n −6n+9

n
}

37. an = {(−1 )
n 1

n
3
}

38. an ={
n

2
n }

39. Prove Theorem 56; that is, use the definition of the limit of a sequence to show that if lim | an | = 0 , then lim an = 0 .
n→∞ n→∞

40. Let a n and bn be sequences such that lim an = L and lim bn = K .


n→∞ n→∞

(a) Show that if a < b for all n, then L ≤ K .


n n

(b) Give an example where L = K .


41. Prove the Squeeze Theorem for sequences: Let a n and bn be such that lim an = L and lim bn = L , and let c be such that
n
n→∞ n→∞

an ≤ cn ≤ bn for all n. Then lim cn = L


n→∞

8.2: Infinite Series


Terms and Concepts
1. Use your own words to describe how sequences and series are related.
2. Use your own words to define a partial sum.

3. Given a series ∑ an m describe the two sequences related to the series that are important.
n=1

4. Use your own words to explain what a geometric series is.

8.E.2 https://math.libretexts.org/@go/page/9979

5. T/F: If a is convergent, then


n ∑ an is also convergent.
n=1

Problems

In Exercises 6-13, a series ∑ an is given.


n=1

(a) Give the first 5 partial sums of the series.


(b) Give a graph of the first 5 terms of a and S on the same axes.
n n

∞ n
(−1)
6. ∑
n
n=1

7. ∑
1

n2
n=1

8. ∑ cos(πn)
n=1

9. ∑ n
n=1

10. ∑
n!
1

n=1

11. ∑
3
1
n

n=1


n
12. ∑ (−
10
9
)
n=1


n
13. ∑ (
10
1
)
n=1

In Exercises 14-19, use Theorem 63 to show the given series diverges.



2

14. ∑
n(n+2)
3n

n=1


n

15. ∑
2

n
2

n=1

16. ∑
10
n!
n

n=1


n 5

17. ∑
5 −n

5 +n
n 5

n=1

∞ n

18. ∑
2 +1

n+1
2
n=1


n
19. ∑ (1 +
1

n
)
n=1

In Exercises 20-29, state whether the given series converges or diverges.


20. ∑
n
1
5

n=1

21. ∑
5
1
n

n=0


n

22. ∑
6

5
n

n=0

23. ∑ n
−4

n=1

8.E.3 https://math.libretexts.org/@go/page/9979


24. ∑ √n
n=1

25. ∑
10

n!
n=1

26. ∑ (
n!
1
+
1

n
)
n=1

27. ∑
2
2
(2x+8)
n=1

28. ∑
1

2n
n=1

29. ∑
2n−1
1

n=1

In Exercises 30-44, a series is given.


(a) Find a formula for S , the n partial sum of the series.n
th

(b) Determine whether the series converges or diverges. If it converges, state what it converges to.

30. ∑
1

4
n

n=0

31. 1 3
+2
3
+3
3 3
+ 4 +. . .

32. ∑ (−1 ) n
n
\)
n=1

33. ∑
5

2
n

n=0

34. ∑ e
−n

n=1

35. 1 − 1

3
+
1

9

1

27
+
1

81
+. . .

36. ∑
n(n+1)
1

n=1

37. ∑
n(n+2)
3

n=1

38. ∑
(2x−1)(2x+1)
1

n=1

39. ∑ ln(
n+1
n
)
n=1


2n+1
40. ∑ 2
n2 (n+1 )
n=1

41. 1

1⋅4
+
2⋅5
1
+
1

3⋅6
+
1

4⋅7
+. . .

42. 2 + ( 1

2
+
1

3
)+(
1

4
+
1

9
)+(
1

8
+
1

27
+. . . )

43. ∑
n −1
1
2

n=2

44. ∑ (sin 1)
n

n=0

45. Break the Harmonic Series into the sum of the odd and even terms:
∞ ∞ ∞


1

n
= ∑
2n−1
1
+ ∑
1

2n
.
n=1 n=1 n=1

The goal is to show that each of the series on the right diverge.

8.E.4 https://math.libretexts.org/@go/page/9979
∞ ∞

(a) Show why ∑


1

2n−1
> ∑
1

2n
. (Compare each n partial sum.)
th

n=1 n=1
∞ ∞

(b) Show why ∑


1

2n−1
<1+ ∑
1

2n
n=1 n=1

(c) Explain why (a) and (b) demonstrate that the series of odd terms is convergent, if, and only if, the series of even terms is also
convergent. (That is, show both converge or both diverge.)
(d) Explain why knowing the Harmonic Series is divergent determines that the even and odd series are also divergent.

46. Show the series ∑


n

(2n−1)(2n+1)
diverges.
n=1

8.3: Integral and Comparison Tests


Terms and Concepts
1. In order to apply the Integral Test to a sequence A , the function a(n) = a must be _____, _____ and _____.
n n

2.T/F: The Integral Test can be used to determine the sum of a convergent series.
3.What test(s) in this section do not work well with factorials?

4. Suppose ∑ an is convergent, and there are sequences b and c such that


n n bn ≤ an ≤ cn for all n. What can be said about the
n=0
∞ ∞

series ∑ bn and ∑ cn ?
n=0 n=0

Problems
In Exercises 5-12, use the Integral Test to determine the convergence of the given series.

5. ∑
2
1
n

n=1

6. ∑
n
1
4

n=1

7. ∑
n +1
2
n

n=1

8. ∑
n ln n
1

n=2

9. ∑
n +1
2
1

n=1

10. ∑
1
2
n(ln n)
n=2

11. ∑
2
n
n

n=1

12. ∑
ln n

n3
n=1

In Exercises 13-22, use the Direct Comparison Test to determine the convergence of the given series; state what series is
used for comparison.

13. ∑ 2
n +3n−5
1

n=1

14. ∑ n
4 +n2 −n
1

n=1

15. ∑
ln n

n
n=1

8.E.5 https://math.libretexts.org/@go/page/9979

16. ∑
n!+n
1

n=1

17. ∑
1

√n2 −1
n=2

18. ∑
1

√n−2
n=1


2
n +n+1
19. ∑
2
n

n=1


n

20. ∑ n
2

5 +10
n=1

21. ∑
n −1
n
2

n=2

22. ∑
n2 ln n
1

n=2

In Exercises 23-32, use the Limit Comparison Test to determine the convergence of the given series; state what series is used
for comparison.

23. ∑ 2
n −3n+5
1

n=1

24. ∑ n
4 −n
1
2

n=1

25. ∑
ln n

n−3
n=4

26. ∑
1

√n2 +n
n=1

27. ∑
n+√n
1

n=1

28. ∑ 2
n−10

n +10n+10
n=1

29. ∑ sin(1/n)
n=1


n+5
30. ∑
n −5
3

n=1


√n+3
31. ∑ 2
n +17
n=1

32. ∑
1

√n+100
n=1

In Exercises 33-40, determine the convergence of the given series. State the test used; more than one test may be
appropriate.

2

33. ∑
n

2
n

n=1

34. ∑
1
3
(2n+5)
n=1

35. ∑
n!

10
n

n=1

8.E.6 https://math.libretexts.org/@go/page/9979

36. ∑
ln n

n!
n=1

37. ∑ n
3 +n
1

n=1

38. ∑
n−2

10n+5
n=1


n

39. ∑
3

n
3

n=1


cos(1/n)
40. ∑
√n
n=1

41. Given that ∑ an converges, state which of the following series converges, may converge, or does not converge.
n=1

an
(a) ∑
n
n=1

(b) ∑ an an+1
n=1

2
(c) ∑ (an )
n=1

(d) ∑ nan
n=1

(e) ∑
an
1

n=1

8.4: Ratio and Root Tests


Terms and Concepts
1. The Ratio Test is not effective when the terms of a sequence only contain ______ functions
2. The Ratio Test is most effective when the terms of a sequence contains ____ and/or _____ functions.
3. What three convergence tests do not work well with terms containing factorials?
4. The Root Test works particularly well on series where each term is _____ to a _____.

Problems
In Exercises 5-14, determine the convergence of the given series using the Ratio Test. If the Ratio Test is inconclusive, state
so and determine the convergence with another test.

5. ∑
2n

n!
n=0


n

6. ∑
5 −3n

4
n

n=0


n

7. ∑
n!10

(2n)!
n=0

∞ n 4

8. ∑
5 +n
n
7 +n
2

n=1

9. ∑
n
1

n=1

10. ∑
3 n +7
1
3

n=1


n

11. ∑
10⋅5
n
7 −3
n=1

8.E.7 https://math.libretexts.org/@go/page/9979

n
12. ∑ n⋅(
3

5
)
n=1

13. ∑
2⋅4⋅6⋅8⋅⋅⋅2n

3⋅6⋅9⋅12⋅⋅⋅3n
n=1

14. ∑
5⋅10⋅15⋅⋅⋅(5n)
n!

n=1

In Exercises 15-24, determine the convergence of the given series using the Root Test. If the Root Test is inconclusive, state
so and determine convergence with another test.
∞ n

15. ∑ (
2n+5

3n+11
)
n=1

∞ n
2

16. ∑ (
0.9 n −n−3
2
n +n+3
)
n=1


n 2

17. ∑
2 n

3
n

n=1

18. ∑
1

nn
n=1


n

19. ∑
2
3
n+1
n 2
n=1


n+7

20. ∑
4

7
n

n=1


2

21. ∑ (
n −n

n2 +n
)
n=1

∞ 2

22. ∑ (
1

n

1
2
n
)
n=1

23. ∑
1
2
(ln n)
n=1


2

24. ∑
(ln n)
n
n

n=1

In Exercises 25-34, determine the convergence of the given series. State the test used; more than one test may be
appropriate.

2
n +4n−2
25. ∑ 3
n +4 n −3n+7
2

n=1


4 n

26. ∑
n 4

n!
n=1


2

27. ∑
3 +n
n
n

n=1


n

28. ∑
3

n
n

n=1

29. ∑
n

√n2 +4n+1
n=1

30. ∑
n!n!n!

(3n)!
n=1

31. ∑
ln n
1

n=1

8.E.8 https://math.libretexts.org/@go/page/9979

32. ∑ (
n+2

n+1
)
n=1


3

33. ∑
n

(ln n)
n

n=1

34. ∑ (
1

n

1

n+2
)
n=1

8.5: Alternating Series and Absolute Convergence


Terms and Concepts

1. Why is ∑ sin n not an alternating series?


n=1

2. A series ∑ (−1 ) an
n
converges when a is _____ _____ and
n lim an = _____.
n→∞
n=1

∞ ∞

3. Give an example of a series when ∑ an converges but ∑ | an | does not.


n=0 n=0

4. The sum of a _____ convergent series can be changed by rearranging the order of its terms.

Problems

In Exercises 5-20, an alternating series ∑ an is given.


n=i

(a) Determine if the series converges or diverges.


(b) Determine if ∑ | an | converges or diverges.


n=0

(c) If ∑ an converges, determine if the convergence is conditional or absolute.


n=0

∞ n+1
(−1)
5. ∑
n
2

n=1

∞ n+1
(−1)
6. ∑
√n!
n=1


n+5
7. ∑ (−1 )
n

3n−5
n=0

8. ∑ (−1 )
n 2n
2
n
n=1

9. ∑ (−1 )
n+1
2
3n+5

n −3n+1
n=0

∞ n
(−1)
10. ∑
ln n+1
n=1

11. ∑ (−1 )
n

ln n
n

n=2

∞ n+1
(−1)
12. ∑
1+3+5+...+(2n−1)
n=1

13. ∑ cos(πn)
n=1


sin((n+1/2)π)
14. ∑
n ln n
n=1


n
15. ∑ (−
2

3
)
n=0

8.E.9 https://math.libretexts.org/@go/page/9979

16. ∑ (−e)
−n

n=0

∞ n 2
(−1) n
17. ∑
n!
n=0


2

18. ∑ (−1 ) 2
n −n

n=0

∞ n
(−1)
19. ∑
√n
n=0

∞ 2
(−1000)
20. ∑
n!
n=1

Let S be the n partial sum of a series. In Exercises 21-24, a convergent alternating series is given and a value of n.
n
th

Compute S and S and use these values to find bounds on the sum of the series.
n n+1

∞ n
(−1)
21. ∑
ln(n+1)
, n =5
n=1

∞ n+1
(−1)
22. ∑
n
4
, n =4
n=1

∞ n
(−1)
23. ∑
n!
, n =6
n=0


n
24. ∑ (−
1

2
) , n =9
n=0

In Exercises 25-28, a convergent alternating series is given along with its sum and a value of ϵ. Use Theorem 71 to find n
such that the n partial sum of the series is within ϵ of a sum of the series.
th

∞ n+1
4
(−1)
25. ∑
n
4
=

720
, ϵ = 0.001
n=0

∞ n
(−1)
26. ∑
n!
=
1

e
, ϵ = 0.0001
n=0

∞ n
(−1)
27. ∑
2n+1
=
π

4
, ϵ = 0.001
n=0

∞ n
(−1)
28. ∑
(2n)!
= cos 1,
−8
ϵ = 10
n=0

8.6: Power Series


Terms and Concepts
1. We adopt the convection that x 0
= _____, regardless of the value of x.
2. What is the difference between the radius of convergence and the interval of convergence?
∞ ∞

3. If the radius of convergence of ∑ ax x


n
is 5, what is the radius of convergence of ∑ n ⋅ an x
n−1
?
n=0 n=1

∞ ∞

4. If the radius of convergence of ∑ an x


n
is 5, what is the radius of convergence of n
∑ (= 1 ) an x
n
?
n=0 n=0

Problems
In Exercises 5-8, write out the sum of the first 5 terms of the given power series.

5. ∑ 2 x
n n

n=0

6. ∑
n2
1
x
n

n=1

8.E.10 https://math.libretexts.org/@go/page/9979

7. ∑
1

n!
x
n

n=0

∞ n
(−1)
8. ∑
(2n)!
x
2n

n=0

In Exercises 9-24, a power series is given.


(a) Find the radius of convergence.
(b) Find the interval of convergence.
∞ n+1
(−1)
9. ∑
n!
x
n

n=0

10. ∑ nx
n

n=0

∞ n n
(−1 ) (x−3 )
11. ∑
n
n=1

∞ n
(x+4)
12. ∑
n!
n=0


n

13. ∑
x

2
n

n=0

∞ n n
(−1 ) (x−5 )
14. ∑
10
n

n=0

15. n
∑ 5 (x − 1 )
n

n=0

16. ∑ (−2 ) x
n n

n=0


− n
17. ∑ √n x
n=0

18. ∑
3
n
n
x
n

n=0


n

19. ∑
3

n!
(x − 5 )
n

n=0

20. ∑ (−1 ) n!(x − 10 )


n n

n=0


n

21. ∑
x

n
2

n=1

∞ n
(x+2)
22. ∑
n
3

n=1


n
23. ∑ n! (
x

10
)
n=0

∞ n
x+4
24. ∑ n (
2

4
)
n=0

In Exercises 25-30, a function f (x) = ∑ an x


n
is given.
n=0

(a) Give a power series for f (x) and its interval of convergence.

(b) Give a power series for f ; (x) dx and its interval of convergence.

25. ∑ nx
n

n=0

8.E.11 https://math.libretexts.org/@go/page/9979

n

26. ∑
x

n
n=1


n
27. ∑ (
x

2
)
n=0

28. ∑ (−3x )
n

n=0

∞ n 2n
(−1) x
29. ∑
(2n)!
n=0

∞ n n
(−1) x
30. ∑
n!
n=0

In Exercises 31-36, give the first 5 terms of the series that is a solution to the given differential equation.
31. y ′
= 3y, y(0) = 1

32. y ′
= 5y, y(0) = 5

33. y ′
=y ,
2
y(0) = 1

34. y ′
= y + 1, y(0) = 1

35. y ′′
= −y, y(0) = 0, y (0) = 1

36. y ′′
= 2y, y(0) = 1, y (0) = 1

8.7: Taylor Polynomials


Terms and Concepts
1. What is the difference between a Taylor polynomial and a Maclaurin polynomial?
2. T/F: In general, p n (x) approximates f (x) better and better as n gets larger.
3. For some function f (x), the Maclaurin polynomial of degree 3 is p 4 (x)
2
= 6 + 3x − 4 x
3
+ 5x − 7x
4
. What is p 2 (x) ?
4. For some function f (x), the Maclaurin polynomial of degree 3 is p 4 (x)
2
= 6 + 3x − 4 x
3
+ 5x − 7x
4
. What is f ′′′
?
(0)

Problems
In Exercises 5-12, find the Maclaurin polynomial of degree n for the given function.
5. f (x) = e −x
, n =3 .
6. f (x) = sin x, n =8 .
7. f (x) = x ⋅ e x
, n =5 .
8. f (x) = tan x, n =6 .
9. f (x) = e 2x
, n =4 .
10. f (x) = 1−x
1
, n =4 .

11. f (x) = 1+x


1
, n =4 .
12. f (x) = 1−x
1
, n =7 .
In Exercises 13-20, find the Taylor polynomial of degree n, at x = c , for the given function.
13. f (x) = √−
x, n = 4, c =1

14. f (x) = ln(x + 1), n = 4, c =1

15. f (x) = cos x, n = 6, c = π/4

16. f (x) = sin x, n = 5, c = π/6

17. f (x) = 1

x
, n = 5, c =2

8.E.12 https://math.libretexts.org/@go/page/9979
18. f (x) = 1

x2
, n = 8, c =1

19. f (x) = x +1
2
1
, n = 4, c = −1

20. f (x) = x 2
cos x, n = 2, c = −1

In Exercises 21-24, approximate the function value with the indicated Taylor polynomial and give approximate bounds on
the error.
21. Approximatesin 0.1 with the Maclaurin polynomial of degree 3.
22. Approximate cos 1 with the Maclaurin polynomial of degree 4.
−−
23. Approximate √10 with the Taylor polynomial of degree 2 centered at x = 9 .
24. Approximate ln 1.5 with the Taylor polynomial of degree 3 centered at x = 1 .
Exercises 25-28 ask for an n to be found such that p n (x) approximates f (x) within a certain bound of accuracy.
25. Find n such that the Maclaurin polynomial of degree n of f (x) = e approximates within 0.0001 of the actual value.
x


26. Find n such that the Taylor polynomial of degree n of f (x) = √−
x , centered at x =4 , approximates √3 within 0.0001 of the
actual value.
27. Find n such that the Maclaurin polynomial of degree n of f (x) = cos x approximates cos π/3 within 0.0001 of the actual value.
28. Find n such that the Maclaurin polynomial of degree n of f (x) = sin x approximates cos π within 0.0001 of the actual value.
In Exercises 29-33, find the n term of the indicated Taylor polynomial.
th

29. Find a formula for the n term of the Maclaurin polynomial for f (x) = e .
th x

30. Find a formula for the n term of the Maclaurin polynomial for f (x) = cos x.
th

31. Find a formula for the n term of the Maclaurin polynomial for f (x) =
th 1

1−x
.

32. Find a formula for the n term of the Maclaurin polynomial for f (x) =
th 1

1+x
.
33. Find a formula for the n term of the Maclaurin polynomial for f (x) = ln x .
th

In Exercises 34-36, approximate the solution to the given differential equation with a degree 4 Maclaurin polynomial.
34. y ′
= y, y(0) = 1

35. y ′
= 5y, y(0) = 3

36. y ′
=
2

y
, y(0) = 1

8.8: Taylor Series


Terms and Concepts
1. What is the difference between a Taylor polynomial and a Taylor series?
2 What theorem must we use to show that a function is equal to its Taylor series?

Problems
Key Idea 32 gives the n term of the Taylor series of common functions. In Exercises 3-6, verify the formula given in the
th

Key Idea by finding the first few terms of the Taylor series of the given function and identifying a pattern.
3. f (x) = e x
; c =0

4. f (x) = sin x; c =0

5. f (x) = 1/(1 − x); c =0

6. f (x) = tan −1
x; c =0

In Exercises 7-12, find a formula for the n term of the Taylor series of f (x), centered at c, by finding the coefficients of the
th

first few powers of x and looking for a pattern. (The formulas for several of these are found in Key Idea 32; show work

8.E.13 https://math.libretexts.org/@go/page/9979
verifying these formula.)
7. f (x) = cos x; c = π/2

8. f (x) = 1/x; c =1

9. f (x) = e −x
; c = π/2

10. f (x) = ln(1 + x); c =0

11. f (x) = x/(x + 1); c =1

12. f (x) = sin x; c = π/4

In Exercises 13-16, show that the Taylor series for f (x), as given in Key Idea 32, is equal to f (x) by applying Theorem 77;
that is, show lim R (x) = 0 . n
n→∞

13. f (x) = e x

14. f (x) = sin x


15. f (x) = ln x
16. f (x) = 1/(1 − x) (show equality only on (-1,0))
In Exercises 17-20, use the Taylor series given in Key Idea 32 to verify the given identity.
17. cos(−x) = cos x
18. sin(−x) = − sin x
19. d

dx
(sin x) = cos x

20. d

dx
(cos x) = − sin x

In Exercises 21-24, write out the first 5 terms of the Binomial series with the given k-value.
21. k = 1/2
22. k = −1/2
23. k = 1/3
24. k = 4
In Exercises 25-30, use the Taylor series given in Key Idea 32 to create the Taylor series of the given functions.
25. f (x) = cos(x 2
)

26. f (x) = e −x

27. f (x) = sin(2x + 3)


28. f (x) = tan −1
(x/2)

29. f (x) = e x
sin x (only find the first 4 terms).
30. f (x) = (1 + x ) 1/2
cos x (only find the first 4 terms)
In Exercises 31-32, approximate the value of the given definite integral by using the first 4 nonzero terms of the integrand's
Taylor series.
√π
31. ∫ 0
sin(x ) dx
2

2
π /4 −
32.∫ 0
cos(√x ) dx

8.E: Applications of Sequences and Series (Exercises) is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by
LibreTexts.

8.E.14 https://math.libretexts.org/@go/page/9979
CHAPTER OVERVIEW
9: Curves in the Plane
We have explored functions of the form y = f (x) closely throughout this text. We have explored their limits, their derivatives and
their antiderivatives; we have learned to identify key features of their graphs, such as relative maxima and minima, inflection points
and asymptotes; we have found equations of their tangent lines, the areas between portions of their graphs and the x-axis, and the
volumes of solids generated by revolving portions of their graphs about a horizontal or vertical axis.
Despite all this, the graphs created by functions of the form y = f (x) are limited. Since each x-value can correspond to only 1 y-
value, common shapes like circles cannot be fully described by a function in this form. Fittingly, the “vertical line test” excludes
vertical lines from being functions of x, even though these lines are important in mathematics.
In this chapter we’ll explore new ways of drawing curves in the plane. We’ll still work within the framework of functions, as an
input will still only correspond to one output. However, our new techniques of drawing curves will render the vertical line test
pointless, and allow us to create important – and beautiful – new curves. Once these curves are defined, we’ll apply the concepts of
calculus to them, continuing to find equations of tangent lines and the areas of enclosed regions.
9.1: Conic Sections
9.2: Parametric Equations
9.3: Calculus and Parametric Equations
9.4: Introduction to Polar Coordinates
9.5: Calculus and Polar Functions
9.E: Applications of Curves in a Plane (Exercises)

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 9: Curves in the Plane is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et
al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

1
9.1: Conic Sections
The ancient Greeks recognized that interesting shapes can be formed by intersecting a plane with a double napped cone (i.e., two
identical cones placed tip--to--tip as shown in the following figures). As these shapes are formed as sections of conics, they have
earned the official name "conic sections.''
The three "most interesting'' conic sections are given in the top row of Figure 9.1.1. They are the parabola, the ellipse (which
includes circles) and the hyperbola. In each of these cases, the plane does not intersect the tips of the cones (usually taken to be the
origin).

Figure 9.1.1 : Conic Sections


When the plane does contain the origin, three degenerate sections can be formed as shown the bottom row of Figure 9.1.1: a point,
a line, and crossed lines. We focus here on the nondegenerate cases.
While the above geometric constructs define the conics in an intuitive, visual way, these constructs are not very helpful when trying
to analyze the shapes algebraically or consider them as the graph of a function. It can be shown that all conics can be defined by the
general second--degree equation
2 2
Ax + Bxy + C y + Dx + Ey + F = 0. (9.1.1)

While this algebraic definition has its uses, most find another geometric perspective of the conics more beneficial. Each
nondegenerate conic can be defined as the locus, or set, of points that satisfy a certain distance property. These distance properties
can be used to generate an algebraic formula, allowing us to study each conic as the graph of a function.

Large Parabolas
Definition 40: parabolas
A parabola is the locus of all points equidistant from a point (called a focus) and a line (called the directrix) that does not
contain the focus.

9.1.1 https://math.libretexts.org/@go/page/4206
Figure 9.1.2 : Illustrating the definition of the parabola and establishing an algebraic formula.
Figure 9.1.1 illustrates this definition. The point halfway between the focus and the directrix is the vertex. The line through the
focus, perpendicular to the directrix, is the axis of symmetry, as the portion of the parabola on one side of this line is the mirror--
image of the portion on the opposite side.
The definition leads us to an algebraic formula for the parabola. Let P = (x, y) be a point on a parabola whose focus is at
F = (0, p) and whose directrix is at y = −p . (We'll assume for now that the focus lies on the y - axis; by placing the focus p units

above the x- axis and the directrix p units below this axis, the vertex will be at (0, 0).)
We use the Distance Formula to find the distance d between F and P :
1

−−−−−−−−−−−−−−−
2 2
d1 = √ (x − 0 ) + (y − p ) . (9.1.2)

The distance d from P to the directrix is more straightforward:


2

d2 = y − (−p) = y + p. (9.1.3)

These two distances are equal. Setting d 1 = d2 , we can solve for y in terms of x:
d1 = d2
−−−−−−−−−−−
2 2
√x + (y − p ) = y +p

Now square both sides.


2 2 2
x + (y − p ) = (y + p)

2 2 2 2 2
x +y − 2yp + p =y + 2yp + p
2
x = 4yp

1
2
y = x .
4p

The geometric definition of the parabola has led us to the familiar quadratic function whose graph is a parabola with vertex at the
origin. When we allow the vertex to not be at (0, 0), we get the following standard form of the parabola.

key idea 33: general equation of a parabola


1. Vertical Axis of Symmetry: The equation of the parabola with vertex at (h, k) and directrix y = k − p in standard form is
1
2
y = (x − h ) + k. (9.1.4)
4p

The focus is at (h, k + p) .


2. Horizontal Axis of Symmetry: The equation of the parabola with vertex at (h, k) and directrix x = h − p in standard
form is
1
2
x = (y − k) + h. (9.1.5)
4p

9.1.2 https://math.libretexts.org/@go/page/4206
The focus is at (h + p, k) .
Note: p is not necessarily a positive number.

Example 9.1.1: Finding the equation of a parabola

Give the equation of the parabola with focus at (1, 2) and directrix at y = 3 .

Solution
The vertex is located halfway between the focus and directrix, so (h, k) = (1, 2.5). This gives p = −0.5 . Using Key Idea 33
we have the equation of the parabola as
1 1
2 2
y = (x − 1 ) + 2.5 = − (x − 1 ) + 2.5. (9.1.6)
4(−0.5) 2

The parabola is sketched in Figure 9.1.3.

Figure 9.1.3 : The parabola described in Example 9.1.1

Example 9.1.2: Finding the focus and directrix of a parabola

Find the focus and directrix of the parabola x = 1

8
y
2
−y +1 . The point (7, 12) lies on the graph of this parabola; verify that it
is equidistant from the focus and directrix.

Solution
We need to put the equation of the parabola in its general form. This requires us to complete the square:
1 2
x = y −y +1
8
1 2
= (y − 8y + 8)
8
1
2
= (y − 8y + 16 − 16 + 8)
8

1 2
= ((y − 4 ) − 8)
8
1 2
= (y − 4 ) − 1.
8

Hence the vertex is located at . We have


(−1, 4)
1

8
=
1

4p
, so p =2 . We conclude that the focus is located at (1, 4) and the
directrix is x = −3 . The parabola is graphed in Figure 9.1.4, along with its focus and directrix.

9.1.3 https://math.libretexts.org/@go/page/4206
The point (7, 12) lies on the graph and is 7 − (−3) = 10 units from the directrix. The distance from (7, 12) to the focus is:
−−−−−−−−−−−−−−−−
2 2 −−−
√ (7 − 1 ) + (12 − 4 ) = √100 = 10. (9.1.7)

Indeed, the point on the parabola is equidistant from the focus and directrix.

Figure 9.1.4 : The parabola described in Example 9.1.2 . The distances from a point on the parabola to the focus and directrix is
given.

Reflective Property
One of the fascinating things about the nondegenerate conic sections is their reflective properties. Parabolas have the following
reflective property:

Any ray emanating from the focus that intersects the parabola reflects off along a line perpendicular to the directrix.
This is illustrated in Figure 9.1.5. The following theorem states this more rigorously.

Figure 9.1.5 : Illustrating the parabola's reflective property.

THEOREM 79: REFLECTIVE PROPERTY OF THE PARABOLA


Let P be a point on a parabola. The tangent line to the parabola at P makes equal angles with the following two lines:
1. The line containing P and the focus F , and
2. The line perpendicular to the directrix through P .

Because of this reflective property, paraboloids (the 3D analogue of parabolas) make for useful flashlight reflectors as the light
from the bulb, ideally located at the focus, is reflected along parallel rays. Satellite dishes also have paraboloid shapes. Signals
coming from satellites effectively approach the dish along parallel rays. The dish then focuses these rays at the focus, where the
sensor is located.

9.1.4 https://math.libretexts.org/@go/page/4206
Ellipses
Definition 41: ellipse

An ellipse is the locus of all points whose sum of distances from two fixed points, each a focus of the ellipse, is constant.

An easy way to visualize this construction of an ellipse is to pin both ends of a string to a board. The pins become the foci. Holding
a pencil tight against the string places the pencil on the ellipse; the sum of distances from the pencil to the pins is constant: the
length of the string. See Figure 9.1.6.

Figure 9.1.6 : Illustrating the construction of an ellipse with pins, pencil and string.
We can again find an algebraic equation for an ellipse using this geometric definition. Let the foci be located along the x- axis, c
units from the origin. Let these foci be labeled as F = (−c, 0) and F = (c, 0) . Let P = (x, y) be a point on the ellipse. The sum
1 2

of distances from F to P (d ) and from F to P (d ) is a constant d . That is, d + d = d . Using the Distance Formula, we have
1 1 2 2 1 2

−−−−−−−−−− −−−−−−−−−−
2 2 2 2
√ (x + c ) +y + √ (x − c ) +y = d. (9.1.8)

Using a fair amount of algebra can produce the following equation of an ellipse (note that the equation is an implicitly defined
function; it has to be, as an ellipse fails the Vertical Line Test):
2 2
x y
+ = 1. (9.1.9)
2 2
d d 2
( ) ( ) −c
2 2

−− −−−−
This is not particularly illuminating, but by making the substitution a = d/2 and b = √a 2
− c2 , we can rewrite the above equation
as
2 2
x y
+ = 1. (9.1.10)
2 2
a b

This choice of a and b is not without reason; as shown in Figure 9.1.7, the values of a and b have geometric meaning in the graph
of the ellipse.

9.1.5 https://math.libretexts.org/@go/page/4206
Figure 9.1.7 : Labeling the significant features of an ellipse.
In general, the two foci of an ellipse lie on the major axis of the ellipse, and the midpoint of the segment joining the two foci is the
center. The major axis intersects the ellipse at two points, each of which is a vertex. The line segment through the center and
perpendicular to the major axis is the minor axis. The "constant sum of distances'' that defines the ellipse is the length of the major
axis, i.e., 2a.

Allowing for the shifting of the ellipse gives the following standard equations.

key idea 34: standard equation of the ellipse

The equation of an ellipse centered at (h, k) with major axis of length 2a and minor axis of length 2b in standard form is:
2 2
(x−h) (y−k)
1. Horizontal major axis: a
2
+ 2
= 1.
b
2 2
(x−h) (y−k)
2. Vertical major axis: 2
+
a
2
= 1.
b

The foci lie along the major axis, c units from the center, where c 2
=a
2
−b
2
.

Example 9.1.3: Finding the equation of an ellipse

Find the general equation of the ellipse graphed in Figure 9.1.8.


Solution
The center is located at (−3, 1). The distance from the center to a vertex is 5 units, hence a = 5 . The minor axis seems to have
length 4, so b = 2 . Thus the equation of the ellipse is
2 2
(x + 3) (y − 1)
+ = 1. (9.1.11)
4 25

9.1.6 https://math.libretexts.org/@go/page/4206
Figure 9.1.8 : The ellipse used in Example 9.1.3 .

Example 9.1.4: Graphing an ellipse

Graph the ellipse defined by 4x 2


+ 9y
2
− 8x − 36y = −4 .
Solution
It is simple to graph an ellipse once it is in standard form. In order to put the given equation in standard form, we must
complete the square with both the x and y terms. We first rewrite the equation by regrouping:
2 2 2 2
4x + 9y − 8x − 36y = −4 ⇒ (4 x − 8x) + (9 y − 36y) = −4. (9.1.12)

Now we complete the squares.


2 2
(4 x − 8x) + (9 y − 36y) = −4
2 2
4(x − 2x) + 9(y − 4y) = −4
2 2
4(x − 2x + 1 − 1) + 9(y − 4y + 4 − 4) = −4

2 2
4((x − 1 ) − 1) + 9((y − 2 ) − 4) = −4

2 2
4(x − 1 ) − 4 + 9(y − 2 ) − 36 = −4

2 2
4(x − 1 ) + 9(y − 2 ) = 36
2 2
(x − 1) (y − 2)
+ = 1.
9 4

We see the center of the ellipse is at (1, 2). We have a = 3 and b = 2 ; the major axis is horizontal, so the vertices are located at
−−− − –
(−2, 2) and (4, 2). We find c = √9 − 4 = √5 ≈ 2.24. The foci are located along the major axis, approximately 2.24 units

from the center, at (1 ± 2.24, 2). This is all graphed in Figure 9.1.9.

Figure 9.1.9 : Graphing the ellipse in Example 9.1.4 .

9.1.7 https://math.libretexts.org/@go/page/4206
Eccentricity
When a = b , we have a circle. The general equation becomes
2 2
(x − h) (y − k)
2 2 2
+ =1 ⇒ (x − h ) + (y − k) =a , (9.1.13)
2 2
a a

−−−−−−
the familiar equation of the circle centered at (h, k) with radius a . Since a = b , c = √a 2
−b
2
=0 . The circle has "two'' foci, but
they lie on the same point, the center of the circle.
Consider Figure 9.1.1, where several ellipses are graphed with a = 1 . In (a), we have c = 0 and the ellipse is a circle. As c grows,
the resulting ellipses look less and less circular. A measure of this "noncircularness'' is eccentricity.

Figure 9.1.10 : Understanding the eccentricity of an ellipse.

Definition 42: eccentricity of an ellipse


The eccentricity e of an ellipse is e = c

a
.

The eccentricity of a circle is 0; that is, a circle has no "noncircularness.'' As c approaches a , e approaches 1, giving rise to a very
noncircular ellipse, as seen in Figure 9.1.10d.
It was long assumed that planets had circular orbits. This is known to be incorrect; the orbits are elliptical. Earth has an eccentricity
of 0.0167 -- it has a nearly circular orbit. Mercury's orbit is the most eccentric, with e = 0.2056. (Pluto's eccentricity is greater, at
e = 0.248 , the greatest of all the currently known dwarf planets.) The planet with the most circular orbit is Venus, with

e = 0.0068 . The Earth's moon has an eccentricity of e = 0.0549 , also very circular.

9.1.8 https://math.libretexts.org/@go/page/4206
Reflective Property
The ellipse also possesses an interesting reflective property. Any ray emanating from one focus of an ellipse reflects off the ellipse
along a line through the other focus, as illustrated in Figure 9.1.11. This property is given formally in the following theorem.

Figure 9.1.11 : Illustrating the reflective property of an ellipse.

theorem 80: reflective property of an ellipse


Let P be a point on a ellipse with foci F and F . The tangent line to the ellipse at P makes equal angles with the following
1 2

two lines:
1. The line through F and P , and
1

2. The line through F and P .


2

This reflective property is useful in optics and is the basis of the phenomena experienced in whispering halls.

Hyperbolas
The definition of a hyperbola is very similar to the definition of an ellipse; we essentially just change the word "sum'' to
"difference.''

Definition 43: hyperbola


A hyperbola is the locus of all points where the absolute value of difference of distances from two fixed points, each a focus
of the hyperbola, is constant.

We do not have a convenient way of visualizing the construction of a hyperbola as we did for the ellipse. The geometric definition
does allow us to find an algebraic expression that describes it. It will be useful to define some terms first.
The two foci lie on the transverse axis of the hyperbola; the midpoint of the line segment joining the foci is the center of the
hyperbola. The transverse axis intersects the hyperbola at two points, each a vertex of the hyperbola. The line through the center
and perpendicular to the transverse axis is the conjugate axis. This is illustrated in Figure 9.1.1. It is easy to show that the constant
difference of distances used in the definition of the hyperbola is the distance between the vertices, i.e., 2a.

Figure 9.1.12 : Labeling the significant features of a hyperbola.

9.1.9 https://math.libretexts.org/@go/page/4206
key idea 35 standard equation of a hyperbola
The equation of a hyperbola centered at (h, k) in standard form is:
2 2
(x−h) (y−k)
1. Horizontal Transverse Axis: a2

2
= 1.
b
2 2
(y−k) (x−h)
2. Vertical Transverse Axis: a
2
− 2
= 1.
b

The vertices are located a units from the center and the foci are located c units from the center, where c 2 2
=a
2
+b .

Graphing Hyperbolas
2 y
2
−−−−−−−
Consider the hyperbola x

9

1
= 1 . Solving for y , we find y = ±√x /9 − 1 . As x grows large, the "−1'' part of the equation
2

−−−−
for y becomes less significant and y ≈ ±√x /9 = ±x/3 . That is, as x gets large, the graph of the hyperbola looks very much like
2

the lines y = ±x/3. These lines are asymptotes of the hyperbola, as shown in Figure 9.1.13.

2 2
y
Figure 9.1.13 : Graphing the hyperbola along with its asymptotes, y = ±x/3 .
x
− = 1
9 1

This is a valuable tool in sketching. Given the equation of a hyperbola in general form, draw a rectangle centered at (h, k) with
sides of length 2a parallel to the transverse axis and sides of length 2b parallel to the conjugate axis. (See Figure 9.1.14 for an
example with a horizontal transverse axis.) The diagonals of the rectangle lie on the asymptotes.

Figure 9.1.14 : Using the asymptotes of a hyperbola as a graphing aid.


These lines pass through (h, k). When the transverse axis is horizontal, the slopes are ±b/a ; when the transverse axis is vertical,
their slopes are ±a/b. This gives equations:

9.1.10 https://math.libretexts.org/@go/page/4206
Example 9.1.5: Graphing a hyperbola
2 2
(y−2) (x−1)
Sketch the hyperbola given by 25

4
= 1.

Solution
The hyperbola is centered at (1, 2); a = 5 and b = 2 . In Figure 9.1.15 we draw the prescribed rectangle centered at (1, 2)
along with the asymptotes defined by its diagonals. The hyperbola has a vertical transverse axis, so the vertices are located at
(1, 7) and (1, −3). This is enough to make a good sketch.

Figure 9.1.15 : Graphing the hyperbola in Example 9.1.5 .


−−
We also find the location of the foci: as c
2
=a
2
+b
2
, we have c = √29 ≈ 5.4 . Thus the foci are located at (1, 2 ± 5.4) as
shown in Figure 9.1.15.

Example 9.1.6: Graphing a hyperbola

Sketch the hyperbola given by 9x 2


−y
2
+ 2y = 10.

Solution
We must complete the square to put the equation in general form. (We recognize this as a hyperbola since it is a general
quadratic equation and the x and y terms have opposite signs.)
2 2

2 2
9x −y + 2y = 10
2 2
9x − (y − 2y) = 10

2 2
9x − (y − 2y + 1 − 1) = 10
2 2
9x − ((y − 1 ) − 1) = 10

2 2
9x − (y − 1 ) =9

2
(y − 1)
2
x − =1
9

9.1.11 https://math.libretexts.org/@go/page/4206
Figure 9.1.16 : Graphing the hyperbola in Example 9.1.6 .
We see the hyperbola is centered at (0, 1), with a horizontal transverse axis, where a = 1 and b = 3 . The appropriate rectangle
is sketched in Figure 9.1.16 along with the asymptotes of the hyperbola. The vertices are located at (±1, 1). We have
− −
c = √10 ≈ 3.2 , so the foci are located at (±3.2, 1) as shown in the figure.

Eccentricity
Definition 44: ECCENTRICITY OF A HYPERBOLA

The eccentricity of a hyperbola is e = c

a
.

Note that this is the definition of eccentricity as used for the ellipse. When c is close in value to a (i.e., e ≈ 1 ), the hyperbola is
very narrow (looking almost like crossed lines). Figure 9.1.17 shows hyperbolas centered at the origin with a = 1 . The graph in (a)
has c = 1.05, giving an eccentricity of e = 1.05 , which is close to 1. As c grows larger, the hyperbola widens and begins to look
like parallel lines, as shown in Figure 9.1.17d.

9.1.12 https://math.libretexts.org/@go/page/4206
Figure 9.1.17 : Understanding the eccentricity of a hyperbola.

Reflective Property
Hyperbolas share a similar reflective property with ellipses. However, in the case of a hyperbola, a ray emanating from a focus that
intersects the hyperbola reflects along a line containing the other focus, but moving away from that focus. This is illustrated in
Figure 9.1.19. Hyperbolic mirrors are commonly used in telescopes because of this reflective property. It is stated formally in the
following theorem.

THEOREM 81: REFLECTIVE PROPERTY OF HYPERBOLAS


Let P be a point on a hyperbola with foci F1 and F2 . The tangent line to the hyperbola at P makes equal angles with the
following two lines:
1. The line through F and P , and
1

2. The line through F and P .


2

Location Determination
Determining the location of a known event has many practical uses (locating the epicenter of an earthquake, an airplane crash site,
the position of the person speaking in a large room, etc.).
To determine the location of an earthquake's epicenter, seismologists use trilateration (not to be confused with triangulation). A
seismograph allows one to determine how far away the epicenter was; using three separate readings, the location of the epicenter

9.1.13 https://math.libretexts.org/@go/page/4206
can be approximated.
A key to this method is knowing distances. What if this information is not available? Consider three microphones at positions A , B
and C which all record a noise (a person's voice, an explosion, etc.) created at unknown location D. The microphone does not
"know'' when the sound was created, only when the sound was detected. How can the location be determined in such a situation?

Figure 9.1.19 : Illustrating the reflective property of a hyperbola.


If each location has a clock set to the same time, hyperbolas can be used to determine the location. Suppose the microphone at
position A records the sound at exactly 12:00, location B records the time exactly 1 second later, and location C records the noise
exactly 2 seconds after that. We are interested in the difference of times. Since the speed of sound is approximately 340 m/s, we can
conclude quickly that the sound was created 340 meters closer to position A than position B . If A and B are a known distance
apart (as shown in Figure 9.1.1a), then we can determine a hyperbola on which D must lie.
The "difference of distances'' is 340; this is also the distance between vertices of the hyperbola. So we know 2a = 340 . Positions A
and B lie on the foci, so 2c = 1000. From this we can find b ≈ 470 and can sketch the hyperbola, given in Figure 9.1.19b. We
only care about the side closest to A . (Why?)
We can also find the hyperbola defined by positions B and C . In this case, 2a = 680 as the sound traveled an extra 2 seconds to
get to C . We still have 2c = 1000, centering this hyperbola at (−500, 500). We find b ≈ 367 . This hyperbola is sketched in Figure
9.1.1c. The intersection point of the two graphs is the location of the sound, at approximately (188, −222.5)
.

Figure 9.1.18 : Using hyperbolas in location detection.


This chapter explores curves in the plane, in particular curves that cannot be described by functions of the form y = f (x). In this
section, we learned of ellipses and hyperbolas that are defined implicitly, not explicitly. In the following sections, we will learn
completely new ways of describing curves in the plane, using parametric equations and polar coordinates, then study these curves
using calculus techniques.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

9.1.14 https://math.libretexts.org/@go/page/4206
This page titled 9.1: Conic Sections is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et
al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

9.1.15 https://math.libretexts.org/@go/page/4206
9.2: Parametric Equations
We are familiar with sketching shapes, such as parabolas, by following this basic procedure:

The rectangular equation y = f (x) works well for some shapes like a parabola with a vertical axis of symmetry, but in the
previous section we encountered several shapes that could not be sketched in this manner. (To plot an ellipse using the above
procedure, we need to plot the "top'' and "bottom'' separately.)
In this section we introduce a new sketching procedure:

Here, x and y are found separately but then plotted together. This leads us to a definition.

Definition 45 Parametric Equations and Curves


Let f and g be continuous functions on an interval I . The set of all points (x, y) = (f (t), g(t)) in the Cartesian plane, as t
varies over I , is the graph of the parametric equations x = f (t) and y = g(t) , where t is the parameter. A curve is a graph
along with the parametric equations that define it.

This is a formal definition of the word curve. When a curve lies in a plane (such as the Cartesian plane), it is often referred to as a
plane curve. Examples will help us understand the concepts introduced in the definition.

Example 9.2.1: Plotting parametric functions

Plot the graph of the parametric equations x = t , y = t + 1 for t in [−2, 2].


2

Solution
We plot the graphs of parametric equations in much the same manner as we plotted graphs of functions like y = f (x): we
make a table of values, plot points, then connect these points with a "reasonable'' looking curve. Figure 9.20(a) shows such a
table of values; note how we have 3 columns.
The points (x, y) from the table are plotted in Figure 9.20(b). The points have been connected with a smooth curve. Each point
has been labeled with its corresponding t -value. These values, along with the two arrows along the curve, are used to indicate
the orientation of the graph. This information helps us determine the direction in which the graph is "moving.''

9.2.1 https://math.libretexts.org/@go/page/4207
Figure 9.20: A table of values of the parametric equations in Example 9.2.1 along with a sketch of their graph.

We often use the letter t as the parameter as we often regard t as representing time. Certainly there are many contexts in which the
parameter is not time, but it can be helpful to think in terms of time as one makes sense of parametric plots and their orientation
(for instance, "At time t = 0 the position is (1, 2) and at time t = 3 the position is (5, 1).'').

Example 9.2.2: Plotting parametric functions

Sketch the graph of the parametric equations x = cos 2


t , y = cos t + 1 for t in [0, π].
Solution
We again start by making a table of values in Figure 9.21(a), then plot the points (x, y) on the Cartesian plane in Figure
9.21(b).

Figure 9.21: A table of values of the parametric equations in Example 9.2.2 along with a sketch of their graph.
It is not difficult to show that the curves in Examples 9.2.1 and 9.2.2 are portions of the same parabola. While the parabola is
the same, the curves are different. In Example 9.2.1, if we let t vary over all real numbers, we'd obtain the entire parabola. In
this example, letting t vary over all real numbers would still produce the same graph; this portion of the parabola would be
traced, and re--traced, infinitely. The orientation shown in Figure 9.21 shows the orientation on [0, π], but this orientation is
reversed on [π, 2π].

9.2.2 https://math.libretexts.org/@go/page/4207
These examples begin to illustrate the powerful nature of parametric equations. Their graphs are far more diverse than the
graphs of functions produced by "y = f (x)" functions.

Technology Note: Most graphing utilities can graph functions given in parametric form. Often the word "parametric'' is
abbreviated as "PAR'' or "PARAM'' in the options. The user usually needs to determine the graphing window (i.e, the minimum and
maximum x- and y -values), along with the values of t that are to be plotted. The user is often prompted to give a t minimum, a t
maximum, and a "t -step'' or "Δt.'' Graphing utilities effectively plot parametric functions just as we've shown here: they plots lots
of points. A smaller t -step plots more points, making for a smoother graph (but may take longer). In Figure 9.20, the t -step is 1; in
Figure 9.21, the t -step is π/4.
One nice feature of parametric equations is that their graphs are easy to shift. While this is not too difficult in the "y = f (x)"
context, the resulting function can look rather messy. (Plus, to shift to the right by two, we replace x with x − 2 , which is counter--
intuitive.) The following example demonstrates this.

Example 9.2.3: Shifting the graph of parametric functions

Sketch the graph of the parametric equations x = t 2


+t ,y =t2
−t . Find new parametric equations that shift this graph to the
right 3 places and down 2.
Solution
The graph of the parametric equations is given in Figure 9.22 (a). It is a parabola with a axis of symmetry along the line y = x ;
the vertex is at (0, 0).
In order to shift the graph to the right 3 units, we need to increase the x-value by 3 for every point. The straightforward way to
accomplish this is simply to add 3 to the function defining x: x = t + t + 3 . To shift the graph down by 2 units, we wish to
2

decrease each y -value by 2, so we subtract 2 from the function defining y : y = t − t − 2 . Thus our parametric equations for
2

the shifted graph are x = t + t + 3 , y = t − t − 2 . This is graphed in Figure 9.22 (b). Notice how the vertex is now at
2 2

(3, −2).

Figure 9.22: Illustrating how to shift graphs in 9.2.3.

Because the x- and y -values of a graph are determined independently, the graphs of parametric functions often possess features not
seen on "y = f (x)" type graphs. The next example demonstrates how such graphs can arrive at the same point more than once.

9.2.3 https://math.libretexts.org/@go/page/4207
Example 9.2.4: Graphs that cross themselves

Plot the parametric functions x =t


3 2
− 5t + 3t + 11 and y =t
2
− 2t + 3 and determine the t -values where the graph
crosses itself.
Solution
Using the methods developed in this section, we again plot points and graph the parametric equations as shown in Figure 9.23.
It appears that the graph crosses itself at the point (2, 6), but we'll need to analytically determine this.

Figure 9.23: A graph of the parametric equations from Example 9.2.4.


We are looking for two different values, say, s and t , where x(s) = x(t) and y(s) = y(t) . That is, the x-values are the same
precisely when the y -values are the same. This gives us a system of 2 equations with 2 unknowns:
3 2 3 2
s − 5s + 3s + 11 = t − 5t + 3t + 11
(9.2.1)
2 2
s − 2s + 3 = t − 2t + 3

Solving this system is not trivial but involves only algebra. Using the quadratic formula, one can solve for t in the second
− −−−−−−− −
equation and find that t = 1 ± √s − 2s + 1 . This can be substituted into the first equation, revealing that the graph crosses
2

itself at t = −1 and t = 3 . We confirm our result by computing x(−1) = x(3) = 2 and y(−1) = y(3) = 6 .

Converting between Rectangular and Parametric Equations


It is sometimes useful to rewrite equations in rectangular form (i.e., y = f (x)) into parametric form, and vice--versa. Converting
from rectangular to parametric can be very simple: given y = f (x), the parametric equations x = t , y = f (t) produce the same
graph. As an example, given y = x , the parametric equations x = t , y = t produce the familiar parabola. However, other
2 2

parametrizations can be used. The following example demonstrates one possible alternative.

Example 9.2.5: Converting from rectangular to parametric


dy
Consider y = x . Find parametric equations x = f (t) , y = g(t) for the parabola where t =
2

dx
. That is, t =a corresponds to
the point on the graph whose tangent line has slope a .
Solution
dy
We start by computing : y = 2x . Thus we set t = 2x . We can solve for x and find x = t/2 . Knowing that y = x , we have
dx
′ 2

y = t /4 . Thus parametric equations for the parabola y = x are


2 2

2
x = t/2 y = t /4. (9.2.2)

To find the point where the tangent line has a slope of −2, we set t = −2 . This gives the point (−1, 1). We can verify that the
slope of the line tangent to the curve at this point indeed has a slope of −2.

We sometimes chose the parameter to accurately model physical behavior.

9.2.4 https://math.libretexts.org/@go/page/4207
Example 9.2.6: Converting from rectangular to parametric

An object is fired from a height of 0ft and lands 6 seconds later, 192ft away. Assuming ideal projectile motion, the height, in
feet, of the object can be described by h(x) = −x /64 + 3x , where x is the distance in feet from the initial location. (Thus
2

h(0) = h(192) = 0 ft.) Find parametric equations x = f (t) , y = g(t) for the path of the projectile where x is the horizontal

distance the object has traveled at time t (in seconds) and y is the height at time t .
Solution
Physics tells us that the horizontal motion of the projectile is linear; that is, the horizontal speed of the projectile is constant.
Since the object travels 192ft in 6s, we deduce that the object is moving horizontally at a rate of 32ft/s, giving the equation
x = 32t . As y = −x /64 + 3x , we find y = −16 t + 96t . We can quickly verify that y = −32 ft/s , the acceleration due to
2 2 ′′′ 2

gravity, and that the projectile reaches its maximum at t = 3 , halfway along its path.
These parametric equations make certain determinations about the object's location easy: 2 seconds into the flight the object is
at the point (x(2), y(2)) = (64, 128). That is, it has traveled horizontally 64ft and is at a height of 128ft, as shown in Figure
9.24.

Figure 9.24: Graphing projectile motion in Example 9.2.6

It is sometimes necessary to convert given parametric equations into rectangular form. This can be decidedly more difficult, as
some "simple'' looking parametric equations can have very "complicated'' rectangular equations. This conversion is often referred
to as "eliminating the parameter,'' as we are looking for a relationship between x and y that does not involve the parameter t .

Example 9.2.7: Eliminating the parameter

Find a rectangular equation for the curve described by


2
1 t
x = and y = . (9.2.3)
2 2
t +1 t +1

Solution
There is not a set way to eliminate a parameter. One method is to solve for t in one equation and then substitute that value in
the second. We use that technique here, then show a second, simpler method.
−−−−−−
Starting with x = 1/(t 2
+ 1) , solve for t : t = ±√1/x − 1 . Substitute this value for t in the equation for y :
2
t
y =
2
t +1

1/x − 1
=
1/x − 1 + 1

1/x − 1
=
1/x

1
=( − 1) ⋅ x
x

= 1 − x.

9.2.5 https://math.libretexts.org/@go/page/4207
Figure 9.25: Graphing parametric and rectangular equations for a graph in Example 9.2.7

Thus y = 1 − x . One may have recognized this earlier by manipulating the equation for y :
2
t 1
y = =1− = 1 − x. (9.2.4)
t2 + 1 t2 + 1

This is a shortcut that is very specific to this problem; sometimes shortcuts exist and are worth looking for.
We should be careful to limit the domain of the function y = 1 − x . The parametric equations limit x to values in (0, 1], thus
to produce the same graph we should limit the domain of y = 1 − x to the same.
The graphs of these functions is given in Figure 9.25. The portion of the graph defined by the parametric equations is given in
a thick line; the graph defined by y = 1 − x with unrestricted domain is given in a thin line.

Example 9.2.8: Eliminating the parameter

Eliminate the parameter in x = 4 cos t + 3 , y = 2 sin t + 1


Solution
We should not try to solve for t in this situation as the resulting algebra/trig would be messy. Rather, we solve for cos t and
sin t in each equation, respectively. This gives

x −3 y −1
cos t = and sin t = . (9.2.5)
4 2

The Pythagorean Theorem gives cos 2 2


t + sin t =1 , so:
2 2
cos t + sin t =1
2 2
x −3 y −1
( ) +( ) =1
4 2
2 2
(x − 3) (y − 1)
+ =1
16 4

Figure 9.26: Graphing the parametric equations x = 4 cos t + 3 , y = 2 sin t + 1 in Example 9.2.8

9.2.6 https://math.libretexts.org/@go/page/4207
This final equation should look familiar -- it is the equation of an ellipse! Figure 9.26 plots the parametric equations,
demonstrating that the graph is indeed of an ellipse with a horizontal major axis and center at (3, 1).

The Pythagorean Theorem can also be used to identify parametric equations for hyperbolas. We give the parametric equations for
ellipses and hyperbolas in the following Key Idea.

KEY IDEA 36 PARAMETRIC EQUATIONS OF ELLIPSES AND HYPERBOLAS


The parametric equations
x = a cos t + h, y = b sin t + k (9.2.6)

define an ellipse with horizontal axis of length 2a and vertical axis of length 2b, centered at (h, k).
The parametric equations
x = a tan t + h, y = ±b sec t + k (9.2.7)

define a hyperbola with vertical transverse axis centered at (h, k), and
x = ±a sec t + h, y = b tan t + k (9.2.8)

defines a hyperbola with horizontal transverse axis. Each has asymptotes at y = ±b/a(x − h) + k .

Special Curves
Figure 9.27 gives a small gallery of "interesting'' and "famous'' curves along with parametric equations that produce them.
Interested readers can begin learning more about these curves through internet searches.
One might note a feature shared by two of these graphs: "sharp corners,'' or cusps. We have seen graphs with cusps before and
determined that such functions are not differentiable at these points. This leads us to a definition.

9.2.7 https://math.libretexts.org/@go/page/4207
Figure 9.27: A gallery of interesting planar curves.

Definition 46 SMOOTH
A curve C defined by x = f (t) , y = g(t) is smooth on an interval I if f and g are continuous on I and not simultaneously 0
′ ′

(except possibly at the endpoints of I ). A curve is piecewise smooth on I if I can be partitioned into subintervals where C is
smooth on each subinterval.

Consider the astroid, given by x = cos3


t , y = sin3
t . Taking derivatives, we have:
′ 2 ′ 2
x = −3 cos t sin t and y = 3 sin t cos t. (9.2.9)

It is clear that each is 0 when t = 0, π/2, π, …. Thus the astroid is not smooth at these points, corresponding to the cusps seen in
the figure. We demonstrate this once more.

Example 9.2.9: Determine where a curve is not smooth

Let a curve C be defined by the parametric equations x =t


3
− 12t + 17 and y =t
2
− 4t + 8 . Determine the points, if any,
where it is not smooth.
Solution
We begin by taking derivatives.
′ 2 ′
x = 3t − 12, y = 2t − 4. (9.2.10)

We set each equal to 0:

9.2.8 https://math.libretexts.org/@go/page/4207
′ 2
x = 0 ⇒ 3t − 12 = 0 ⇒ t = ±2
(9.2.11)

y = 0 ⇒ 2t − 4 = 0 ⇒ t = 2

We see at t = 2 both x and y are 0; thus C is not smooth at t = 2 , corresponding to the point (1, 4). The curve is graphed in
′ ′

Figure 9.28, illustrating the cusp at (1, 4).

Figure 9.28: Graphing the curve in Example 9.2.9; note it is not smooth at (1, 4).

If a curve is not smooth at t = t , it means that x (t ) = y (t ) = 0 as defined. This, in turn, means that rate of change of x (and
0

0

0

y ) is 0; that is, at that instant, neither x nor y is changing. If the parametric equations describe the path of some object, this means

the object is at rest at t . An object at rest can make a "sharp'' change in direction, whereas moving objects tend to change direction
0

in a "smooth'' fashion.
One should be careful to note that a "sharp corner'' does not have to occur when a curve is not smooth. For instance, one can verify
that x = t , y = t produce the familiar y = x parabola. However, in this parametrization, the curve is not smooth. A particle
3 6 2

traveling along the parabola according to the given parametric equations comes to rest at t = 0 , though no sharp point is created.\\
Our previous experience with cusps taught us that a function was not differentiable at a cusp. This can lead us to wonder about
derivatives in the context of parametric equations and the application of other calculus concepts. Given a curve defined
parametrically, how do we find the slopes of tangent lines? Can we determine concavity? We explore these concepts and more in
the next section.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 9.2: Parametric Equations is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory
Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

9.2.9 https://math.libretexts.org/@go/page/4207
9.3: Calculus and Parametric Equations
The previous section defined curves based on parametric equations. In this section we'll employ the techniques of calculus to study
these curves. We are still interested in lines tangent to points on a curve. They describe how the y -values are changing with respect
to the x-values, they are useful in making approximations, and they indicate instantaneous direction of travel.
dy
The slope of the tangent line is still , and the Chain Rule allows us to calculate this in the context of parametric equations. If
dx

x = f (t) and y = g(t) , the Chain Rule states that

dy dy dx
= ⋅ . (9.3.1)
dt dx dt

dy
Solving for dx
, we get

dy dy dx g (t)
= / = , (9.3.2)

dx dt dt f (t)

provided that f ′
(t) ≠ 0 . This is important so we label it a Key Idea.
dy
key idea 37 Finding dx
with Parametric Equations.

Let x = f (t) and y = g(t) , where f and g are differentiable on some open interval I and f ′
(t) ≠ 0 on I . Then

dy g (t)
= . (9.3.3)

dx f (t)

We use this to define the tangent line.

Definition 47 Tangent and Normal Lines


Let a curve C be parametrized by x = f (t) and y = g(t) , where f and g are differentiable functions on some interval I
containing t = t . The tangent line to C at t = t is the line through (f (t
0 0 0 ), g(t0 ))with slope m = g (t )/f (t ) , provided

0

0

f (t ) ≠ 0 .

0

The normal line to C at t = t is the line through (f (t


0 0 ), g(t0 )) with slope m = −f ′ ′
(t0 )/ g (t0 ) , provided g ′
(t0 ) ≠ 0 .

The definition leaves two special cases to consider. When the tangent line is horizontal, the normal line is undefined by the above
definition as g (t ) = 0 . Likewise, when the normal line is horizontal, the tangent line is undefined. It seems reasonable that these

0

lines be defined (one can draw a line tangent to the "right side'' of a circle, for instance), so we add the following to the above
definition.
1. If the tangent line at t = t has a slope of 0, the normal line to C at t = t is the line x = f (t ) .
0 0 0

2. If the normal line at t = t has a slope of 0, the tangent line to C at t = t is the line x = f (t ) .
0 0 0

Example 9.3.1: Tangent and Normal Lines to Curves

Let x = 5t 2
− 6t + 4 and y = t 2
+ 6t − 1 , and let C be the curve defined by these equations.
1. Find the equations of the tangent and normal lines to C at t = 3 .
2. Find where C has vertical and horizontal tangent lines.
Solution
1. We start by computing f ′
(t) = 10t − 6 and g ′
(t) = 2t + 6 . Thus
dy 2t + 6
= . (9.3.4)
dx 10t − 6

9.3.1 https://math.libretexts.org/@go/page/4208
dy
Make note of something that might seem unusual: dx
is a function of t , not x. Just as points on the curve are found in
terms of t , so are the slopes of the tangent lines.

The point on C at t = 3 is (31, 26). The slope of the tangent line is m = 1/2 and the slope of the normal line is m = −2 .
Thus,

∙ the equation of the tangent line is y = (x − 31) + 26 , and


1

∙ the equation of the normal line is y = −2(x − 31) + 26 .

This is illustrated in Figure 9.29.


dy
2. To find where C has a horizontal tangent line, we set dx
=0 and solve for t . In this case, this amounts to setting g ′
(t) = 0

and solving for t (and making sure that f (t) ≠ 0 ). ′


g (t) = 0 ⇒ 2t + 6 = 0 ⇒ t = −3. (9.3.5)

The point on C corresponding to t = −3 is (67, −10); the tangent line at that point is horizontal (hence with equation
y = −10 ).


f (t)
To find where C has a vertical tangent line, we find where it has a horizontal normal line, and set − g ′ (t)
=0 . This amounts
to setting f ′
(t) = 0 and solving for t (and making sure that g ′
(t) ≠ 0 ).

f (t) = 0 ⇒ 10t − 6 = 0 ⇒ t = 0.6. (9.3.6)

The point on C corresponding to t = 0.6 is (2.2, 2.96). The tangent line at that point is x = 2.2.

The points where the tangent lines are vertical and horizontal are indicated on the graph in Figure 9.29.

Figure 9.29: Graphing tangent and normal lines in Example 9.3.1

Example 9.3.2: Tangent and Normal Lines to a Circle

1. Find where the unit circle, defined by x = cos t and y = sin t on [0, 2π], has vertical and horizontal tangent lines.
2. Find the equation of the normal line at t = t . 0

Solution
1. We compute the derivative following Key Idea 37:

dy g (t) cos t
= =− . (9.3.7)

dx f (t) sin t

The derivative is 0 when cos t = 0 ; that is, when t = π/2, 3π/2 . These are the points (0, 1) and (0, −1) on the circle.

9.3.2 https://math.libretexts.org/@go/page/4208
The normal line is horizontal (and hence, the tangent line is vertical) when sin t = 0 ; that is, when t = 0, π, 2π ,
corresponding to the points (−1, 0) and (0, 1) on the circle. These results should make intuitive sense.
sin t0
2. The slope of the normal line at t = t is m = = tan t . This normal line goes through the point (cos t
0
cos t0
0 0, ,
sin t0 )

giving the line


sin t0
y = (x − cos t0 ) + sin t0
cos t0

= (tan t0 )x,

as long as cos t ≠ 0 . It is an important fact to recognize that the normal lines to a circle pass through its center, as
0

illustrated in Figure 9.30. Stated in another way, any line that passes through the center of a circle intersects the circle at
right angles.

Figure 9.30: Illustrating how a circle's normal lines pass through its center.

Example 9.3.3: Tangent lines when is not defined


dy

dx

Find the equation of the tangent line to the astroid x = cos 3


t , y = sin 3
t at t = 0 , shown in Figure 9.31.
Solution
We start by finding x (t) and y
′ ′
(t) :
′ 2 ′ 2
x (t) = −3 sin t cos t, y (t) = 3 cos t sin t. (9.3.8)

dy
Note that both of these are 0 at t = 0 ; the curve is not smooth at t = 0 forming a cusp on the graph. Evaluating dx
at this point
returns the indeterminate form of "0/0''.
We can, however, examine the slopes of tangent lines near t = 0 , and take the limit as t → 0 .
′ 2
y (t) 3 cos t sin t
lim = lim (We can cancel as t ≠ 0.)

t→0 x (t) t→0 −3 sin t cos2 t

sin t
= lim −
t→0 cos t

= 0.

dy
We have accomplished something significant. When the derivative dx
returns an indeterminate form at t = t , we can define
0

dy
its value by setting it to be lim
dx
, if that limit exists. This allows us to find slopes of tangent lines at cusps, which can be very
t→t0

beneficial.

9.3.3 https://math.libretexts.org/@go/page/4208
Figure 9.31: A graph of an astroid.
We found the slope of the tangent line at t = 0 to be 0; therefore the tangent line is y = 0 , the x- axis.

Concavity
2
d y
We continue to analyze curves in the plane by considering their concavity; that is, we are interested in dx2
, "the second derivative
dy
of y with respect to x.'' To find this, we need to find the derivative of dx
with respect to x; that is,
2
d y d dy
= [ ], (9.3.9)
2
dx dx dx

dy
but recall that dx
is a function of t , not x, making this computation not straightforward.
dy
To make the upcoming notation a bit simpler, let h(t) = dx
. We want d

dx
[h(t)] ; that is, we want dh

dx
. We again appeal to the Chain
Rule. Note:

dh dh dx dh dh dx
= ⋅ ⇒ = / . (9.3.10)
dt dx dt dx dt dt

2
d y dy
In words, to find dx
2
, we first take the derivative of dx
with respect to t , then divide by x (t). We restate this as a Key Idea.

key idea 38 Finding with Parametric Equations


d y

2
dx

Let x = f (t) and y = g(t) be twice differentiable functions on an open interval I , where f ′
(t) ≠ 0 on I . Then
2
d y d dy dx d dy

= [ ]/ = [ ] /f (t). (9.3.11)
2
dx dt dx dt dt dx

Examples will help us understand this Key Idea.

Example 9.3.4: Concavity of Plane Curves

Let x = 5t 2
− 6t + 4 and y =t
2
+ 6t − 1 as in Example 9.3.1. Determine the t - intervals on which the graph is concave
up/down.
Solution
2
d y
Concavity is determined by the second derivative of y with respect to x, dx
2
, so we compute that here following Key Idea 38.
dy
In Example 9.3.1, we found dx
=
2t+6

10t−6
and f ′
(t) = 10t − 6 . So:

9.3.4 https://math.libretexts.org/@go/page/4208
2
d y d 2t + 6
= [ ] /(10t − 6)
2
dx dt 10t − 6

72
=− /(10t − 6)
2
(10t − 6)

72
=−
(10t − 6)3

9
=−
3
(5t − 3)

Figure 9.32: Graphing the parametric equations in Example 9.3.4 to demonstrate concavity.
2 2
d y d y
The graph of the parametric functions is concave up when > 0 and concave down when
dx
2
dx
2
<0 . We determine the
intervals when the second derivative is greater/less than 0 by first finding when it is 0 or undefined.
2
d y
As the numerator of − 9
3
is never 0, dx2
≠0 for all t . It is undefined when 5t − 3 = 0 ; that is, when t = 3/5 . Following
(5t−3)

the work established in Section 3.4, we look at values of t greater/less than 3/5 on a number line:

Reviewing Example 9.3.1, we see that when t = 3/5 = 0.6 , the graph of the parametric equations has a vertical tangent line.
This point is also a point of inflection for the graph, illustrated in Figure 9.32.

Example 9.3.5: Concavity of Plane Curves

Find the points of inflection of the graph of the parametric equations x = √t , y = sin t , for 0 ≤ t ≤ 16 .
Solution
2
dy d y
We need to compute dx
and dx2
.

dy y (t) cos t
= = = 2 √t cos t. (9.3.12)
dx x′ (t) 1/(2 √t)

d dy
2 [ ]
d y dt dx cos t/ √t − 2 √t sin t
= = = 2 cos t − 4t sin t. (9.3.13)
2 ′
dx x (t) 1/(2 √t)

2
d y
The points of inflection are found by setting dx
2
=0 . This is not trivial, as equations that mix polynomials and trigonometric
functions generally do not have "nice'' solutions.
In Figure 9.33(a) we see a plot of the second derivative. It shows that it has zeros at approximately
t = 0.5, 3.5, 6.5, 9.5, 12.5 and 16. These approximations are not very good, made only by looking at the graph. Newton's

Method provides more accurate approximations. Accurate to 2 decimal places, we have:


t = 0.65, 3.29, 6.36, 9.48, 12.61 and 15.74. (9.3.14)

9.3.5 https://math.libretexts.org/@go/page/4208
The corresponding points have been plotted on the graph of the parametric equations in Figure 9.33(b). Note how most occur
near the x- axis, but not exactly on the axis.

Figure 9.33: In (a), a graph of , showing where it is approximately 0. In (b), graph of the parametric equations in Example
d y

2
dx

9.3.5 along with points of inflection.

Arc Length
We continue our study of the features of the graphs of parametric equations by computing their arc length. Recall in Section 7.4 we
found the arc length of the graph of a function, from x = a to x = b , to be
−−−−−−−−−
b 2
dy
L =∫ √1 + ( ) dx. (9.3.15)
a
dx

We can use this equation and convert it to the parametric equation context. Letting x = f (t) and y = g(t) , we know that
dy

dx
′ ′
= g (t)/ f (t) . It will also be useful to calculate the differential of x:
1

dx = f (t)dt ⇒ dt = ⋅ dx. (9.3.16)

f (t)

Starting with the arc length formula above, consider:


−−−−−−−−−
b 2
dy
L =∫ √1 + ( ) dx
a
dx

−−−−−−−−−
b ′ 2
g (t)
=∫ √1 + dx.
′ 2
a f (t)

′ 2
Factor out the f (t) :
b
−−−−−−−−−−− 1
′ 2 ′ 2
=∫ √ f (t) + g (t) ⋅ dx

a f (t)

=dt

t2 −−−−−−−−−−−
′ 2 ′ 2
=∫ √ f (t) + g (t) dt.
t1

9.3.6 https://math.libretexts.org/@go/page/4208
Note the new bounds (no longer "x" bounds, but "t " bounds). They are found by finding t1 and t2 such that a = f (t1 ) and
b = f (t ) . This formula is important, so we restate it as a theorem.
2

theorem 82 arc length of parametric curves


Let x = f (t) and y = g(t) be parametric equations with f and g continuous on some open interval I containing t and t on
′ ′
1 2

which the graph traces itself only once. The arc length of the graph, from t = t to t = t , is 1 2

t2 −−−−−−−−−−−
′ 2 ′ 2
L =∫ √ f (t) + g (t) dt. (9.3.17)
t1

As before, these integrals are often not easy to compute. We start with a simple example, then give another where we approximate
the solution.

Example 9.3.6: Arc Length of a Circle

Find the arc length of the circle parametrized by x = 3 cos t, y = 3 sin t on [0, 3π/2].
Solution
By direct application of Theorem 82, we have
3π/2 −−−−−−−−−−−−−−−−−
2 2
L =∫ √ (−3 sin t) + (3 cos t) dt.
0

Apply the Pythagorean Theorem.

3π/2

=∫ 3 dt
0

3π/2

= 3t = 9π/2.

0

This should make sense; we know from geometry that the circumference of a circle with radius 3 is 6π; since we are finding
the arc length of 3/4 of a circle, the arc length is 3/4 ⋅ 6π = 9π/2.

Example 9.3.7: Arc Length of a Parametric Curve

The graph of the parametric equations x = t(t 2


− 1) ,y =t 2
−1 crosses itself as shown in Figure 9.34, forming a "teardrop.''
Find the arc length of the teardrop.
Solution
We can see by the parametrizations of x and y that when t = ±1 , x = 0 and y = 0 . This means we'll integrate from t = −1 to
t = 1 . Applying Theorem 82, we have

1 −−−−−−−−−−−−−−
2 2 2
L =∫ √ (3 t − 1) + (2t) dt
−1

1
− −−−−−−−− −
√ 4 2
=∫ 9 t − 2 t + 1 dt.
−1

Unfortunately, the integrand does not have an antiderivative expressible by elementary functions. We turn to numerical
integration to approximate its value. Using 4 subintervals, Simpson's Rule approximates the value of the integral as 2.65051.
Using a computer, more subintervals are easy to employ, and n = 20 gives a value of 2.71559. Increasing n shows that this
value is stable and a good approximation of the actual value.

9.3.7 https://math.libretexts.org/@go/page/4208
Figure 9.34: A graph of the parametric equations in Example 9.3.7, where the arc length of the teardrop is calculated.

Surface Area of a Solid of Revolution


Related to the formula for finding arc length is the formula for finding surface area. We can adapt the formula found in Key Idea 28
from Section 7.4 in a similar way as done to produce the formula for arc length done before.

key idea 39 Surface Area of a Solid of Revolution


Consider the graph of the parametric equations x = f (t) and y = g(t) , where f and g are continuous on an open interval
′ ′
I

containing t and t on which the graph does not cross itself.


1 2

1. The surface area of the solid formed by revolving the graph about the x- axis is (where g(t) ≥ 0 on [t 1, ):
t2 ]

t2 −−−−−−−−−−−
′ 2 ′ 2
Surface Area = 2π ∫ g(t)√ f (t) + g (t) dt. (9.3.18)
t1

2. The surface area of the solid formed by revolving the graph about the y - axis is (where f (t) ≥ 0 on [t
1, ):
t2 ]

t2 −−−−−−−−−−−
′ 2 ′ 2
Surface Area = 2π ∫ f (t)√ f (t) + g (t) dt. (9.3.19)
t1

Example 9.3.8: Surface Area of a Solid of Revolution

Consider the teardrop shape formed by the parametric equations x = t(t − 1) , y = t 2 2


−1 as seen in Example 9.3.7. Find the
surface area if this shape is rotated about the x- axis, as shown in Figure 9.3.8.

Figure 9.35: Rotating a teardrop shape about the x-axis in Example 9.3.8
Solution
The teardrop shape is formed between t = −1 and t = 1 . Using Key Idea 39, we see we need for g(t) ≥ 0 on [−1, 1], and this
is not the case. To fix this, we simplify replace g(t) with −g(t) , which flips the whole graph about the x- axis (and does not
change the surface area of the resulting solid). The surface area is:

9.3.8 https://math.libretexts.org/@go/page/4208
1 −−−−−−−−−−−−−−
2 2 2 2
Area S = 2π ∫ (1 − t )√ (3 t − 1) + (2t) dt
−1

1
− −−−−−−−− −
2 √ 4 2
= 2π ∫ (1 − t ) 9 t − 2 t + 1 dt.
−1

Once again we arrive at an integral that we cannot compute in terms of elementary functions. Using Simpson's Rule with
n = 20 , we find the area to be S = 9.44 . Using larger values of n shows this is accurate to 2 places after the decimal.

After defining a new way of creating curves in the plane, in this section we have applied calculus techniques to the parametric
equation defining these curves to study their properties. In the next section, we define another way of forming curves in the plane.
To do so, we create a new coordinate system, called polar coordinates, that identifies points in the plane in a manner different than
from measuring distances from the y - and x- axes.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 9.3: Calculus and Parametric Equations is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by
Gregory Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available
upon request.

9.3.9 https://math.libretexts.org/@go/page/4208
9.4: Introduction to Polar Coordinates
We are generally introduced to the idea of graphing curves by relating x-values to y -values through a function f . That is, we set
y = f (x), and plot lots of point pairs (x, y) to get a good notion of how the curve looks. This method is useful but has limitations,

not least of which is that curves that "fail the vertical line test'' cannot be graphed without using multiple functions.
The previous two sections introduced and studied a new way of plotting points in the x, y-plane. Using parametric equations, x and
y values are computed independently and then plotted together. This method allows us to graph an extraordinary range of curves.

This section introduces yet another way to plot points in the plane: using polar coordinates.

Polar Coordinates
Start with a point O in the plane called the pole (we will always identify this point with the origin). From the pole, draw a ray,
called the initial ray (we will always draw this ray horizontally, identifying it with the positive x-axis). A point P in the plane is
¯
¯¯¯¯¯¯
¯
determined by the distance r that P is from O, and the angle θ formed between the initial ray and the segment OP (measured
counter-clockwise). We record the distance and angle as an ordered pair (r, θ) . To avoid confusion with rectangular coordinates, we
will denote polar coordinates with the letter P , as in P (r, θ). This is illustrated in Figure 9.4.1.

Figure 9.4.1 : Illustrating polar coordinates.


Practice will make this process more clear.

Example 9.4.1: Plotting Polar Coordinates

Plot the following polar coordinates:


A = P (1, π/4) B = P (1.5, π) C = P (2, −π/3) D = P (−1, π/4) (9.4.1)

Solution

Figure 9.4.2 : Plotting polar points in Example 9.4.1


To aid in the drawing, a polar grid is provided at the bottom of this page. To place the point A , go out 1 unit along the initial
ray (putting you on the inner circle shown on the grid), then rotate counter-clockwise π/4 radians (or 45 ). Alternately, one can

consider the rotation first: think about the ray from O that forms an angle of π/4 with the initial ray, then move out 1 unit
along this ray (again placing you on the inner circle of the grid).
To plot B , go out 1.5 units along the initial ray and rotate π radians (180 ). ∘

To plot C , go out 2 units along the initial ray then rotate clockwise π/3 radians, as the angle given is negative.

9.4.1 https://math.libretexts.org/@go/page/4209
To plot D, move along the initial ray "−1" units -- in other words, "back up'' 1 unit, then rotate counter-clockwise by π/4. The
results are given in Figure 9.4.2.

Consider the following two points: A = P (1, π) and B = P (−1, 0) . To locate A , go out 1 unit on the initial ray then rotate π
radians; to locate B , go out −1 units on the initial ray and don't rotate. One should see that A and B are located at the same point
in the plane. We can also consider C = P (1, 3π), or D = P (1, −π) ; all four of these points share the same location.
This ability to identify a point in the plane with multiple polar coordinates is both a "blessing'' and a "curse.'' We will see that it is
beneficial as we can plot beautiful functions that intersect themselves (much like we saw with parametric functions). The
unfortunate part of this is that it can be difficult to determine when this happens. We'll explore this more later in this section.

Polar to Rectangular Conversion


It is useful to recognize both the rectangular (or, Cartesian) coordinates of a point in the plane and its polar coordinates. Figure
9.4.3 shows a point P in the plane with rectangular coordinates (x, y) and polar coordinates P (r, θ). Using trigonometry, we can

make the identities given in the following Key Idea.

Figure 9.4.3 : Converting between rectangular and polar coordinates.

KEY IDEA 40 Converting Between Rectangular and Polar Coordinates


Given the polar point P (r, θ), the rectangular coordinates are determined by

x = r cos θ y = r sin θ. (9.4.2)

Given the rectangular coordinates (x, y), the polar coordinates are determined by
2 2 2
y
r =x +y tan θ = . (9.4.3)
x

Example 9.4.2: Converting Between Polar and Rectangular Coordinates

1. Convert the polar coordinates P (2, 2π/3) and P (−1, 5π/4) to rectangular coordinates.
2. Convert the rectangular coordinates (1, 2) and (−1, 1) to polar coordinates.
1. (a) We start with P (2, 2π/3). Using Key Idea 40, we have

x = 2 cos(2π/3) = −1 y = 2 sin(2π/3) = √3. (9.4.4)


So the rectangular coordinates are (−1, √3) ≈ (−1, 1.732) .

(b) The polar point P (−1, 5π/4) is converted to rectangular with:


– –
x = −1 cos(5π/4) = √2/2 y = −1 sin(5π/4) = √2/2. (9.4.5)

– –
So the rectangular coordinates are (√2/2, √2/2) ≈ (0.707, 0.707).

These points are plotted in Figure 9.4.4 (a). The rectangular coordinate system is drawn lightly under the polar coordinate
system so that the relationship between the two can be seen.
2. (a) To convert the rectangular point (1, 2) to polar coordinates, we use the Key Idea to form the following two equations:

9.4.2 https://math.libretexts.org/@go/page/4209
2 2 2
2
1 +2 =r tan θ = . (9.4.6)
1


The first equation tells us that r = √5 . Using the inverse tangent function, we find
−1 ∘
tan θ = 2 ⇒ θ = tan 2 ≈ 1.11 ≈ 63.43 . (9.4.7)


Thus polar coordinates of (1, 2) are P (√5, 1.11).

(b)To convert (−1, 1) to polar coordinates, we form the equations

2 2 2
1
(−1 ) +1 =r tan θ = . (9.4.8)
−1


Thus r = √2 . We need to be careful in computing θ : using the inverse tangent function, we have
−1 ∘
tan θ = −1 ⇒ θ = tan (−1) = −π/4 = −45 . (9.4.9)

This is not the angle we desire. The range of tan x is (−π/2, π/2); that is, it returns angles that lie in the 1 and 4
−1 st th

quadrants. To find locations in the 2 and 3 quadrants, add π to the result of tan x. So π + (−π/4) puts the angle at
nd rd −1


3π/4. Thus the polar point is P (√2, 3π/4).

An alternate method is to use the angle θ given by arctangent, but change the sign of r. Thus we could also refer to (−1, 1)

as\\ P (−√2, −π/4).
These points are plotted in Figure 9.4.4 (b). The polar system is drawn lightly under the rectangular grid with rays to
demonstrate the angles used.

Figure 9.4.4 : Plotting rectangular and polar points in Example 9.4.2

Polar Functions and Polar Graphs


Defining a new coordinate system allows us to create a new kind of function, a polar function. Rectangular coordinates lent
themselves well to creating functions that related x and y , such as y = x . Polar coordinates allow us to create functions that relate
2

r and θ . Normally these functions look like r = f (θ) , although we can create functions of the form θ = f (r) . The following

examples introduce us to this concept.

Example 9.4.3: Introduction to Graphing Polar Functions

Describe the graphs of the following polar functions.


1. r = 1.5
2. θ = π/4

9.4.3 https://math.libretexts.org/@go/page/4209
Solution
1. The equation r = 1.5 describes all points that are 1.5 units from the pole; as the angle is not specified, any θ is allowable.
All points 1.5 units from the pole describes a circle of radius 1.5.

We can consider the rectangular equivalent of this equation; using r = x + y , we see that 1.5 = x + y , which we
2 2 2 2 2 2

recognize as the equation of a circle centered at (0, 0) with radius 1.5. This is sketched in Figure 9.4.5.
2. The equation θ = π/4 describes all points such that the line through them and the pole make an angle of π/4 with the
initial ray. As the radius r is not specified, it can be any value (even negative). Thus θ = π/4 describes the line through the
pole that makes an angle of π/4 = 45 with the initial ray.

We can again consider the rectangular equivalent of this equation. Combine tan θ = y/x and θ = π/4 :
tan π/4 = y/x ⇒ x tan π/4 = y ⇒ y = x. (9.4.10)

This graph is also plotted in Figure 9.4.5.

Figure 9.4.5 : Plotting standard polar plots.

The basic rectangular equations of the form x = h and y = k create vertical and horizontal lines, respectively; the basic polar
equations r = h and θ = α create circles and lines through the pole, respectively. With this as a foundation, we can create more
complicated polar functions of the form r = f (θ) . The input is an angle; the output is a length, how far in the direction of the angle
to go out.
We sketch these functions much like we sketch rectangular and parametric functions: we plot lots of points and "connect the dots''
with curves. We demonstrate this in the following example.

Example 9.4.4: Sketching Polar Functions

Sketch the polar function r = 1 + cos θ on [0, 2π] by plotting points.


Solution
A common question when sketching curves by plotting points is "Which points should I plot?'' With rectangular equations, we
often choose "easy'' values -- integers, then added more if needed. When plotting polar equations, start with the "common''
angles -- multiples of π/6 and π/4. Figure 9.4.6 gives a table of just a few values of θ in [0, π].
Consider the point P (0, 2) determined by the first line of the table. The angle is 0 radians -- we do not rotate from the initial

ray -- then we go out 2 units from the pole. When θ = π/6 , r = 1.866 (actually, it is 1 + √3/2 ); so rotate by π/6 radians and
go out 1.866 units.
The graph shown uses more points, connected with straight lines. (The points on the graph that correspond to points in the
table are signified with larger dots.) Such a sketch is likely good enough to give one an idea of what the graph looks like.

9.4.4 https://math.libretexts.org/@go/page/4209
Figure 9.4.6 : Graphing a polar function in Example 9.4.4 by plotting points.

Technology Note
Plotting functions in this way can be tedious, just as it was with rectangular functions. To obtain very accurate graphs,
technology is a great aid. Most graphing calculators can plot polar functions; in the menu, set the plotting mode to something
like polar or POL , depending on one's calculator. As with plotting parametric functions, the viewing "window'' no longer
determines the x-values that are plotted, so additional information needs to be provided. Often with the "window'' settings are
the settings for the beginning and ending θ values (often called θ and θ ) as well as the θ
min max -- that is, how far apart the θ
step

values are spaced. The smaller the θ step value, the more accurate the graph (which also increases plotting time). Using
technology, we graphed the polar function r = 1 + cos θ from Example 9.4.4 in Figure 9.4.7.

Figure 9.4.7 : Using technology to graph a polar function.

Example 9.4.5: Sketching Polar Functions

Sketch the polar function r = cos(2θ) on [0, 2π] by plotting points.


Solution
We start by making a table of cos(2θ) evaluated at common angles θ , as shown in Figure 9.4.8. These points are then plotted
in Figure 9.4.9 (a). This particular graph "moves'' around quite a bit and one can easily forget which points should be
connected to each other. To help us with this, we numbered each point in the table and on the graph.

9.4.5 https://math.libretexts.org/@go/page/4209
Figure 9.4.8 : Tables of points for plotting a polar curve.
Using more points (and the aid of technology) a smoother plot can be made as shown in Figure 9.4.9 (b). This plot is an
example of a rose curve.

Figure 9.4.9 : Polar plots from Example 9.4.5

It is sometimes desirable to refer to a graph via a polar equation, and other times by a rectangular equation. Therefore it is
necessary to be able to convert between polar and rectangular functions, which we practice in the following example. We will make
frequent use of the identities found in Key Idea 40.

Example 9.4.6: Converting between rectangular and polar equations.

Convert from rectangular to polar.


1. y = x 2

2. xy = 1
Convert from polar to rectangular.
1. r = 2

sin θ−cos θ

2. r = 2 cos θ
Solution
1. Replace y with r sin θ and replace x with r cos θ, giving:
2
y =x
2 2
r sin θ = r cos θ

sin θ
=r
cos2 θ

We have found that r = sin θ/ cos θ = tan θ sec θ . The domain of this polar function is (−π/2, π/2); plot a few points to
2

see how the familiar parabola is traced out by the polar equation.
2. We again replace x and y using the standard identities and work to solve for r:

9.4.6 https://math.libretexts.org/@go/page/4209
xy = 1

r cos θ ⋅ r sin θ = 1

1
2
r =
cos θ sin θ
1
r =
− −−− − −−−
√ cos θ sin θ

This function is valid only when the product of cos θ sin θ is positive. This occurs in the first and third quadrants, meaning
the domain of this polar function is (0, π/2) ∪ (π, 3π/2).

We can rewrite the original rectangular equation xy = 1 as y = 1/x. This is graphed in Figure 9.4.10; note how it only
exists in the first and third quadrants.
3. There is no set way to convert from polar to rectangular; in general, we look to form the products r cos θ and r sin θ , and
then replace these with x and y , respectively. We start in this problem by multiplying both sides by sin θ − cos θ :
2
r =
sin θ − cos θ

r(sin θ − cos θ) = 2

r sin θ − r cos θ = 2. Now replace with y and x:

y −x = 2

y = x + 2.

The original polar equation, r = 2/(sin θ − cos θ) does not easily reveal that its graph is simply a line. However, our
conversion shows that it is. The upcoming gallery of polar curves gives the general equations of lines in polar form.
4. By multiplying both sides by r, we obtain both an r term and an r cos θ term, which we replace with x + y and x,
2 2 2

respectively.
r = 2 cos θ
2
r = 2r cos θ
2 2
x +y = 2x.

We recognize this as a circle; by completing the square we can find its radius and center.
2 2
x − 2x + y =0 (9.4.11)
2 2
(x − 1 ) +y = 1. (9.4.12)

The circle is centered at (1, 0) and has radius 1. The upcoming gallery of polar curves gives the equations of some circles in
polar form; circles with arbitrary centers have a complicated polar equation that we do not consider here.

Figure 9.4.10 : Graphing xy = 1 from Example 9.4.6

9.4.7 https://math.libretexts.org/@go/page/4209
Some curves have very simple polar equations but rather complicated rectangular ones. For instance, the equation r = 1 + cos θ
describes a cardiod (a shape important the sensitivity of microphones, among other things; one is graphed in the gallery in the
Lima\c con section). It's rectangular form is not nearly as simple; it is the implicit equation
4 4 2 2 2 3 2
x +y + 2x y − 2x y − 2x −y = 0. (9.4.13)

The conversion is not "hard,'' but takes several steps, and is left as a problem in the Exercise section.

Gallery of Polar Curves


There are a number of basic and "classic'' polar curves, famous for their beauty and/or applicability to the sciences. This section
ends with a small gallery of some of these graphs. We encourage the reader to understand how these graphs are formed, and to
investigate with technology other types of polar functions.

9.4.8 https://math.libretexts.org/@go/page/4209
Earlier we discussed how each point in the plane does not have a unique representation in polar form. This can be a "good'' thing,
as it allows for the beautiful and interesting curves seen in the preceding gallery. However, it can also be a "bad'' thing, as it can be
difficult to determine where two curves intersect.

Example 9.4.7: Finding points of intersection with polar curves

Determine where the graphs of the polar equations r = 1 + 3 cos θ and r = cos θ intersect.

9.4.9 https://math.libretexts.org/@go/page/4209
Solution
As technology is generally readily available, it is usually a good idea to start with a graph. We have graphed the two functions
in Figure 9.4.11(a); to better discern the intersection points, part (b) of the figure zooms in around the origin.

Figure 9.4.11 : Graphs to help determine the points of intersection of the polar functions given in Example 9.4.7
We start by setting the two functions equal to each other and solving for θ :

1 + 3 cos θ = cos θ

2 cos θ = −1

1
cos θ = −
2
2π 4π
θ = , .
3 3

(There are, of course, infinite solutions to the equation cos θ = −1/2 ; as the lima\c con is traced out once on , we
[0, 2π]

restrict our solutions to this interval.)


We need to analyze this solution. When θ = 2π/3 we obtain the point of intersection that lies in the 4 quadrant. When th

θ = 4π/3 , we get the point of intersection that lies in the 2 quadrant. There is more to say about this second intersection
nd

point, however. The circle defined by r = cos θ is traced out once on [0, π], meaning that this point of intersection occurs while
tracing out the circle a second time. It seems strange to pass by the point once and then recognize it as a point of intersection
only when arriving there a "second time.'' The first time the circle arrives at this point is when θ = π/3 .
It is key to understand that these two points are the same: (cos π/3, π/3)and (cos 4π/3, 4π/3).
To summarize what we have done so far, we have found two points of intersection: when θ = 2π/3 and when θ = 4π/3 .
When referencing the circle r = cos θ , the latter point is better referenced as when θ = π/3 .
There is yet another point of intersection: the pole (or, the origin). We did not recognize this intersection point using our work
above as each graph arrives at the pole at a different θ value.
A graph intersects the pole when r = 0 . Considering the circle r = cos θ , r = 0 when θ = π/2 (and odd multiples thereof, as
the circle is repeatedly traced). The lima\c con intersects the pole when 1 + 3 cos θ = 0 ; this occurs when cos θ = −1/3 , or
for θ = cos (−1/3) . This is a nonstandard angle, approximately θ = 1.9106 = 10\(9.4.12^\circ\). The lima\c con intersects
−1

the pole twice in [0, 2π]; the other angle at which the lima\c con is at the pole is the reflection of the first angle across the x-
axis. That is, θ = 4.3726 = 250.53 . ∘

If all one is concerned with is the (x, y) coordinates at which the graphs intersect, much of the above work is extraneous. We know
they intersect at (0, 0); we might not care at what θ value. Likewise, using θ = 2π/3 and θ = 4π/3 can give us the needed
rectangular coordinates. However, in the next section we apply calculus concepts to polar functions. When computing the area of a
region bounded by polar curves, understanding the nuances of the points of intersection becomes important.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

9.4.10 https://math.libretexts.org/@go/page/4209
This page titled 9.4: Introduction to Polar Coordinates is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by
Gregory Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available
upon request.

9.4.11 https://math.libretexts.org/@go/page/4209
9.5: Calculus and Polar Functions
The previous section defined polar coordinates, leading to polar functions. We investigated plotting these functions and solving a
fundamental question about their graphs, namely, where do two polar graphs intersect?
We now turn our attention to answering other questions, whose solutions require the use of calculus. A basis for much of what is
done in this section is the ability to turn a polar function r = f (θ) into a set of parametric equations. Using the identities
x = r cos θ and y = r sin θ , we can create the parametric equations x = f (θ) cos θ , y = f (θ) sin θ and apply the concepts of

Section 9.3.

Polar Functions and


dy

dx

We are interested in the lines tangent a given graph, regardless of whether that graph is produced by rectangular, parametric, or
dy
polar equations. In each of these contexts, the slope of the tangent line is dx
. Given r = f (θ) , we are generally not concerned with
dy

r = f (θ)

; that describes how fast r changes with respect to θ . Instead, we will use x = f (θ) cos θ , y = f (θ) sin θ to compute dx
.
Using Key Idea 37 we have
dy dy dx
= / . (9.5.1)
dx dθ dθ

Each of the two derivatives on the right hand side of the equality requires the use of the Product Rule. We state the important result
as a Key Idea.

key idea 41 Finding with Polar Functions


dy

dx

Let r = f (θ) be a polar function. With x = f (θ) cos θ and y = f (θ) sin θ ,

dy f (θ) sin θ + f (θ) cos θ
= . (9.5.2)

dx f (θ) cos θ − f (θ) sin θ

dy
Example 9.5.1: Finding dx
with polar functions.

Consider the limacon r = 1 + 2 sin θ on [0, 2π].


1. Find the equations of the tangent and normal lines to the graph at θ = π/4 .
2. Find where the graph has vertical and horizontal tangent lines.
Solution
dy
1. We start by computing dx
. With f ′
(θ) = 2 cos θ , we have

dy 2 cos θ sin θ + cos θ(1 + 2 sin θ)


=
2
dx 2 cos θ − sin θ(1 + 2 sin θ)

cos θ(4 sin θ + 1)


= .
2 2
2(cos θ − sin θ) − sin θ

dy –
When θ = π/4 , = −2 √2 − 1 (this requires a bit of simplification). In rectangular coordinates, the point on the graph at
dx
– –
θ = π/4 is (1 + √2/2, 1 + √2/2) . Thus the rectangular equation of the line tangent to the limacon at θ = π/4 is

– – –
y = (−2 √2 − 1)(x − (1 + √2/2)) + 1 + √2/2 ≈ −3.83x + 8.24. (9.5.3)

The limacon and the tangent line are graphed in Figure 9.47.
The normal line has the opposite--reciprocal slope as the tangent line, so its equation is
1
y ≈ x + 1.26. (9.5.4)
3.83

9.5.1 https://math.libretexts.org/@go/page/7467
dy dy
2. To find the horizontal lines of tangency, we find where dx
=0 ; thus we find where the numerator of our equation for dx
is
0.
cos θ(4 sin θ + 1) = 0 ⇒ cos θ = 0 or 4 sin θ + 1 = 0. (9.5.5)

On [0, 2π], cos θ = 0 when θ = π/2, 3π/2 .

Setting 4 sin θ + 1 = 0 gives θ = sin (−1/4) ≈ −0.2527 = −14.48 . We want the results in [0, 2π]; we also recognize
−1 ∘

there are two solutions, one in the 3 quadrant and one in the 4 . Using reference angles, we have our two solutions as
rd th

θ = 3.39 and 6.03 radians. The four points we obtained where the limacon has a horizontal tangent line are given in Figure

9.47 with black--filled dots.

dy
To find the vertical lines of tangency, we set the denominator of dx
=0 .
2 2
2(cos θ − sin θ) − sin θ = 0.

2 2
Convert the cos θ term to 1 − sin θ:
2 2
2(1 − sin θ − sin θ) − sin θ = 0

2
4 sin θ + sin θ − 1 = 0.

Recognize this as a quadratic in the variable sin θ. Using the quadratic formula, we have
−−
−1 ± √33
sin θ = .
8

−1+√33 −1−√33
We solve sin θ = 8
and sin θ = 8
:
−− −−
−1 + √33 −1 − √33
sin θ = sin θ =
8 8
−− −−
−1
−1 + √33 −1
−1 − √33
θ = sin ( ) θ = sin ( )
8 8

θ = 0.6399 θ = −1.0030

In each of the solutions above, we only get one of the possible two solutions as sin −1
x only returns solutions in
[−π/2, π/2] , the 4 and 1 quadrants. Again using reference angles, we have:
th st

−−
−1 + √33
sin θ = ⇒ θ = 0.6399, 3.7815 radians (9.5.6)
8

and
−−
−1 − √33
sin θ = ⇒ θ = 4.1446, 5.2802 radians. (9.5.7)
8

These points are also shown in Figure 9.47 with white--filled dots.

9.5.2 https://math.libretexts.org/@go/page/7467
Figure 9.47: The limacon in Example 9.5.1 with its tangent line at θ = π/4 and points of vertical and horizontal tangency.

When the graph of the polar function r = f (θ) intersects the pole, it means that f (α) = 0 for some angle α . Thus the formula for
dy

dx
in such instances is very simple, reducing simply to
dy
= tan α. (9.5.8)
dx

This equation makes an interesting point. It tells us the slope of the tangent line at the pole is tan α ; some of our previous work
(see, for instance, Example 9.4.3) shows us that the line through the pole with slope tan α has polar equation θ = α . Thus when a
polar graph touches the pole at θ = α , the equation of the tangent line at the pole is θ = α .

Example 9.5.2: Finding tangent lines at the pole.

Let r = 1 + 2 sin θ , a limacon. Find the equations of the lines tangent to the graph at the pole.

Figure 9.48: Graphing the tangent lines at the pole in Example 9.5.2.
Solution
We need to know when r = 0 .
1 + 2 sin θ = 0

sin θ = −1/2

7π 11π
θ = , .
6 6

Thus the equations of the tangent lines, in polar, are θ = 7π/6 and θ = 11π/6. In rectangular form, the tangent lines are
y = tan(7π/6)x and y = tan(11π/6)x. The full limacon con can be seen in Figure 9.47; we zoom in on the tangent lines in

Figure 9.48.
Note: Recall that the area of a sector of a circle with radius r subtended by an angle θ is A = 1

2
2
θr .

9.5.3 https://math.libretexts.org/@go/page/7467
Area
When using rectangular coordinates, the equations x = h and y = k defined vertical and horizontal lines, respectively, and
combinations of these lines create rectangles (hence the name "rectangular coordinates''). It is then somewhat natural to use
rectangles to approximate area as we did when learning about the definite integral.
When using polar coordinates, the equations θ = α and r = c form lines through the origin and circles centered at the origin,
respectively, and combinations of these curves form sectors of circles. It is then somewhat natural to calculate the area of regions
defined by polar functions by first approximating with sectors of circles.
Consider Figure 9.49 (a) where a region defined by r = f (θ) on [α, β] is given. (Note how the "sides'' of the region are the lines
θ = α and θ = β , whereas in rectangular coordinates the "sides'' of regions were often the vertical lines x = a and x = b .)

Figure 9.49: Computing the area of a polar region.


Partition the interval [α, β] into n equally spaced subintervals as α = θ < θ < ⋯ < θ
1 = β . The length of each subinterval
2 n+1

is Δθ = (β − α)/n , representing a small change in angle. The area of the region defined by the i subinterval [θ , θ ] can be
th
i i+1

approximated with a sector of a circle with radius f (c ), for some c in [θ , θ ]. The area of this sector is f (c ) Δθ . This is
i i i i+1
1

2
i
2

shown in part (b) of the figure, where [α, β] has been divided into 4 subintervals. We approximate the area of the whole region by
summing the areas of all sectors:
n
1
2
Area ≈ ∑ f (ci ) Δθ. (9.5.9)
2
i=1

This is a Riemann sum. By taking the limit of the sum as n → ∞ , we find the exact area of the region in the form of a definite
integral.

THEOREM 83 AREA OF A POLAR REGION


Let f be continuous and non-negative on [α, β], where 0 ≤ β − α ≤ 2π . The area A of the region bounded by the curve
r = f (θ) and the lines θ = α and θ = β is
β β
1 1
2 2
A = ∫ f (θ) dθ = ∫ r dθ (9.5.10)
2 α
2 α

The theorem states that 0 ≤ β − α ≤ 2π . This ensures that region does not overlap itself, which would give a result that does not
correspond directly to the area.

9.5.4 https://math.libretexts.org/@go/page/7467
Example 9.5.3: Area of a polar region

Find the area of the circle defined by r = cos θ . (Recall this circle has radius 1/2.)
Solution
This is a direct application of Theorem 83. The circle is traced out on [0, π], leading to the integral
π
1
2
Area = ∫ cos θ dθ
2 0

π
1 1 + cos(2θ)
= ∫ dθ
2 0
2
π
1 1 ∣
= (θ + sin(2θ))∣
4 2 ∣
0

1
= π.
4

Of course, we already knew the area of a circle with radius 1/2 . We did this example to demonstrate that the area formula is
correct.

Note: Example 9.5.3 requires the use of the integral ∫ cos θ dθ . This is handled well by using the power reducing formula as
2

found in the back of this text. Due to the nature of the area formula, integrating cos θ and sin θ is required often. We offer here
2 2

these indefinite integrals as a time--saving measure.

2
1 1
∫ cos θ dθ = θ+ sin(2θ) + C (9.5.11)
2 4

2
1 1
∫ sin θ dθ = θ− sin(2θ) + C (9.5.12)
2 4

Example 9.5.4: Area of a polar region

Find the area of the cardiod r = 1 + cos θ bound between θ = π/6 and θ = π/3 , as shown in Figure 9.50.

Figure 9.50: Finding the area of the shaded region of a cardiod in Example 9.5.4
Solution
This is again a direct application of Theorem 83.
π/3
1 2
Area = ∫ (1 + cos θ) dθ
2 π/6

π/3
1 2
= ∫ (1 + 2 cos θ + cos θ) dθ
2 π/6

π/3
1 1 1 ∣
= (θ + 2 sin θ + θ+ sin(2θ)) ∣
2 2 4 ∣
π/6

1 –
= (π + 4 √3 − 4) ≈ 0.7587.
8

9.5.5 https://math.libretexts.org/@go/page/7467
Area Between Curves
Our study of area in the context of rectangular functions led naturally to finding area bounded between curves. We consider the
same in the context of polar functions. \index{polar!functions!area between curves}
Consider the shaded region shown in Figure 9.51. We can find the area of this region by computing the area bounded by
r = f (θ) and subtracting the area bounded by r = f (θ) on [α, β]. Thus
2 2 1 1

β β β
1 1 1
2 2 2 2
Area = ∫ r dθ − ∫ r dθ = ∫ (r −r ) dθ. (9.5.13)
2 1 2 1
2 α
2 α
2 α

Figure 9.51: Illustrating area bound between two polar curves.

KEY IDEA 42 area between polar curves

The area A of the region bounded by r 1 = f1 (θ) and r 2 = f2 (θ) , θ = α and θ = β , where f 1 (θ) ≤ f2 (θ) on [α, β], is
β
1
2 2
A = ∫ (r −r ) dθ. (9.5.14)
2 1
2 α

Example 9.5.5: Area between polar curves

Find the area bounded between the curves r = 1 + cos θ and r = 3 cos θ , as shown in Figure 9.52.

Figure 9.52: Finding the area between polar curves in Example 9.5.5.
Solution
We need to find the points of intersection between these two functions. Setting them equal to each other, we find:
1 + cos θ = 3 cos θ

cos θ = 1/2

θ = ±π/3

Thus we integrate 1

2
2
((3 cos θ) − (1 + cos θ) )
2
on [−π/3, π/3].

9.5.6 https://math.libretexts.org/@go/page/7467
π/3
1
2 2
Area = ∫ ((3 cos θ) − (1 + cos θ) ) dθ
2 −π/3

π/3
1
2
= ∫ (8 cos θ − 2 cos θ − 1) dθ
2 −π/3

π/3

= (2 sin(2θ) − 2 sin θ + 3θ)∣

−π/3

= 2π.

Amazingly enough, the area between these curves has a "nice'' value

Example 9.5.6: Area defined by polar curves

Find the area bounded between the polar curves r = 1 and r = 2 cos(2θ) , as shown in Figure 9.53 (a).
Solution
We need to find the point of intersection between the two curves. Setting the two functions equal to each other, we have
1
2 cos(2θ) = 1 ⇒ cos(2θ) = ⇒ 2θ = π/3 ⇒ θ = π/6. (9.5.15)
2

Figure 9.53: Graphing the region bounded by the functions in Example 9.5.6
In part (b) of the figure, we zoom in on the region and note that it is not really bounded between two polar curves, but rather by
two polar curves, along with θ = 0 . The dashed line breaks the region into its component parts. Below the dashed line, the
region is defined by r = 1 , θ = 0 and θ = π/6 . (Note: the dashed line lies on the line θ = π/6 .) Above the dashed line the
region is bounded by r = 2 cos(2θ) and θ = π/6 . Since we have two separate regions, we find the area using two separate
integrals.
Call the area below the dashed line A and the area above the dashed line A . They are determined by the following integrals:
1 2

π/6 π/4
1 1 2
2
A1 = ∫ (1 ) dθ A2 = ∫ (2 cos(2θ)) dθ. (9.5.16)
2 0
2 π/6

(The upper bound of the integral computing A is π/4 as r = 2 cos(2θ) is at the pole when θ = π/4 .)
2


We omit the integration details and let the reader verify that A1 = π/12 and A2 = π/12 − √3/8 ; the total area is

A = π/6 − √3/8 .

9.5.7 https://math.libretexts.org/@go/page/7467
Arc Length
As we have already considered the arc length of curves defined by rectangular and parametric equations, we now consider it in the
context of polar equations. Recall that the arc length L of the graph defined by the parametric equations x = f (t) , y = g(t) on
[a, b] is

b −−−−−−−−−−− b −−−−−−−−−−−
′ 2 ′ 2 ′ 2 ′ 2
L =∫ √ f (t) + g (t) dt = ∫ √ x (t) + y (t) dt. (9.5.17)
a a

Now consider the polar function r = f (θ) . We again use the identities x = f (θ) cos θ and y = f (θ) sin θ to create parametric
dy
equations based on the polar function. We compute x (θ) and y (θ) as done before when computing
′ ′
, then apply Equation dx

9.5.17.

The expression x (θ)


′ 2 ′
+ y (θ)
2
can be simplified a great deal; we leave this as an exercise and state that
′ 2 ′ 2 ′ 2 2
x (θ) + y (θ) = f (θ) + f (θ) . (9.5.18)

This leads us to the arc length formula.

key idea 43 arc length of polar curves


Let r = f (θ) be a polar function with f continuous on an open interval

I containing [α, β], on which the graph traces itself
only once. The arc length L of the graph on [α, β] is
β −−−−−−−−−−− β −−−−−−−−
′ 2 2 ′ 2 2
L =∫ √ f (θ) + f (θ) dθ = ∫ √ (r ) +r dθ. (9.5.19)
α α

Example 9.5.7: Arc length of a limacon

Find the arc length of the limacon r = 1 + 2 sin t .

Solution
With r = 1 + 2 sin t , we have r ′
= 2 cos t . The limacon is traced out once on [0, 2π] , giving us our bounds of integration.
Applying Key Idea 43, we have
2π −−−−−−−−−−−−−−−−−−−
2 2
L =∫ √ (2 cos θ) + (1 + 2 sin θ) dθ
0


−−−−−−−−−−−−−−−−−−−−−− −
2 2
=∫ √ 4 cos θ + 4 sin θ + 4 sin θ + 1 dθ


− −−− −− −−
=∫ √ 4 sin θ + 5 dθ
0

≈ 13.3649.

Figure 9.54: The limacon in Example 9.5.7 whose arc length is measured.

9.5.8 https://math.libretexts.org/@go/page/7467
The final integral cannot be solved in terms of elementary functions, so we resorted to a numerical approximation. (Simpson's
Rule, with n = 4 , approximates the value with 13.0608. Using n = 22 gives the value above, which is accurate to 4 places
after the decimal.)

Surface Area
The formula for arc length leads us to a formula for surface area. The following Key Idea is based on Key Idea 39.

KEY IDEA 44 SURFACE AREA OF A SOLID OF REVOLUTION


Consider the graph of the polar equation r = f (θ) , where f is continuous on an open interval containing [α, β] on which the

graph does not cross itself.


1. The surface area of the solid formed by revolving the graph about the initial ray (θ = 0 ) is:
β −−−−−−−−−−−
′ 2 2
Surface Area = 2π ∫ f (θ) sin θ√ f (θ) + f (θ) dθ. (9.5.20)
α

2. The surface area of the solid formed by revolving the graph about the line θ = π/2 is:
β −−−−−−−−−−−
′ 2 2
Surface Area = 2π ∫ f (θ) cos θ√ f (θ) + f (θ) dθ. (9.5.21)
α

Example 9.5.8: Surface area determined by a polar curve

Find the surface area formed by revolving one petal of the rose curve r = cos(2θ) about its central axis (see Figure 9.55.

Solution
We choose, as implied by the figure, to revolve the portion of the curve that lies on [0, π/4] about the initial ray. Using Key
Idea ??? and the fact that f (θ) = −2 sin(2θ) , we have

π/4 −−−−−−−−−−−−−−−−−−−−−−
2 2
Surface Area = 2π ∫ cos(2θ) sin(θ)√ ( − 2 sin(2θ)) + ( cos(2θ)) dθ
0

≈ 1.36707.

The integral is another that cannot be evaluated in terms of elementary functions. Simpson's Rule, with n =4 , approximates
the value at 1.36751.%; with n = 10 , the value is accurate to 4 decimal places.

Figure 9.55: Finding the surface area of a rose-curve petal that is revolved around its central axis.

9.5.9 https://math.libretexts.org/@go/page/7467
This chapter has been about curves in the plane. While there is great mathematics to be discovered in the two dimensions of a
plane, we live in a three dimensional world and hence we should also look to do mathematics in 3D -- that is, in space. The next
chapter begins our exploration into space by introducing the topic of vectors, which are incredibly useful and powerful
mathematical objects.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 9.5: Calculus and Polar Functions is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory
Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

9.5.10 https://math.libretexts.org/@go/page/7467
9.E: Applications of Curves in a Plane (Exercises)
9.1: Conic Sections
Terms and Concepts
1. What is the difference between degenerate and nondegenerate conics?
2. Use your own words to explain what the eccentricity of an ellipse measures.
3. What has the largest eccentricity: an ellipse or a hyperbola?
4. Explain why the following is true: “If the coefficient of the x term in the equation of an ellipse in standard form is smaller than
2

the coefficient of the y term, then the ellipse has a horizontal major axis.”
2

5. Explain how one can quickly look at the equation of a hyperbola in standard form and determine whether the transverse axis is
horizontal or vertical.

Problems
In Exercises 6-13, find the equation of the parabola defined by the given information. Sketch the parabola.
6. Focus: (3, 2); directrix: y = 1
7. Focus: (−1, −4); directrix: y = 2
8. Focus: (1, 5); directrix: x = 3
9. Focus: (1/4, 0); directrix: x = −1/4
10. Focus: (1, 1); vertex: (1, 2)
11. Focus: (−3, 0); vertex: (0, 0)
12. Vertex: (0, 0); directrix: y = −1/16
13. Vertex: (2, 3); directrix: x = 4
In Exercises 14-15, the equation of a parabola and a point on its graph are given. Find the focus and directrix of the
parabola, and verify that the given point is equidistant from the focus and directrix.
14. y = 1

4
2
x , P = (2, 1)

15. x = 1

8
(y − 2 )
2
+ 3, P = (11, 10)

In Exercises 16-17, sketch the ellipse defined by the given equation. Label the center, foci and vertices.
2 2
(x−1) (y−2)
16. 3
+
5
=1

17. 1

25
x
2
+
1

9
(y + 3 )
2
=1

In Exercises 18-19, find the equation of the ellipse shown in the graph. Give the location of the foci and the eccentricity of
the ellipse.
18.

9.E.1 https://math.libretexts.org/@go/page/4771
19.

In Exercises 20-23, find the equation of the ellipse defined by the given information. Sketch the ellipse.
20. Foci: (±2, 0); vertices: (±3, 0)
21. Foci: (−1, 3) and (5, 3); vertices: (−3, 3) and (7, 3)
22. Foci: (2, ±2); vertices: (2, ±7)
23. Focus: (−1, 5); vertex: (−1, −4); center: (−1, 1)
In Exercises 24-27, write the equation of the given ellipse in standard form.
24. x 2
− 2x + 2 y
2
− 8y = −7

25. 5x 2
+ 3y
2
= 15

26. 3x 2
+ 2y
2
− 12y + 6 = 0

27. x 2
+y
2
− 4x − 4y + 4 = 0

2 2
(x−1) (y−3)
28. Consider the ellipse given by +
4
=1 .
12

(a) Verify that the foci are located at (1, 3 ± 2√2)
– –
(b) The point P = (2, 6) and P = (1 + √2, 3 + √6) ≈ (2.414, 5.449) lie on the ellipse. Verify that the sum of distances from
1 2

each point to the foci is the same.


In Exercises 29-32, find the equation of the hyperbola shown in the graph.
29.

30.

9.E.2 https://math.libretexts.org/@go/page/4771
31.

32.

In Exercises 33-34, sketch the hyperbola defined by the given equation. Label the center and foci.
2 2
(x−1) (y+2)
33. 16

9
=1

2
(x+1)
34. (y − 4) 2

25
=1

In Exercises 35-38, find the equation of the hyperbola defined by the given information. Sketch the hyperbola.
35. Foci: (±3, 0); vertices: (±2, 0)
36. Foci: (0, ±3); vertices: (0, ±2)
37. Foci: (−2, 3) and (8, 3) vertices: (−1, 3) and (7, 3)
38. Foci: (3, −2) and (3, 8) vertices: (3, 0) and (3, 6)
In Exercises 39-42, write the equation of the hyperbola in standard form.
39. 3x 2
− 4y
2
= 12

40. 3x 2
−y
2
+ 2y = 10

41. x 2
− 10 y
2
+ 40y = 30

42. (4y − x)(4y + x) = 4


43. Johannes Kepler discovered that the planets of our solar system have elliptical orbits with the Sun at one focus. The Earth's
elliptical orbit is used as a standard unit of distance; the distance from the center of Earth's elliptical orbit to one vertex is 1
Astronomical Unit, or A.U.
The following table gives information about the orbits of three planets.

−−−−−
(a) In an ellipse, knowing c 2 2
=a
2
−b and e = c/a allows us to find b in terms of a and e. Show b = a√1 − e . 2

2
2
y
(b) For each planet, find equations of their elliptical orbit of the form x

a
2
+ 2
=1 . (This places the center at (0,0), but the Sun is
b

in a different location for each planet.)


(c) Shift the equations so that the Sun lies at the origin. Plot the three elliptical orbits.
44. A loud sound is recorded at three stations that lie on a line as shown in the figure below. Station A recorded the sound 1 second
after Station B, and Station C recorded the sound 3 seconds after B. Using the speed of sound as 340m/s, determine the location of

9.E.3 https://math.libretexts.org/@go/page/4771
the sound's origination.

9.2: Parametric Equations


Terms and Concepts
1. T/F: When sketching the graph of parametric equations, the x and y values are found separately, then plotted together.
2. The direction in which a graph is "moving" is called the _____ of the graph.
3. An Equation written as y = f (x) is written in ____ form.
4. Create parametric equations x = f (t), y = g(t) and sketch their graph. Explain any interesting features of your graph based on
the functions f and g .

Problems
In Exercises 5-8, sketch the graph of the given parametric equations by hand, making a table of points to plot. Be sure to
indicate the orientation of the graph.
5. x = t2
+ t, y = 1 −t ,
2
−3 ≤ t ≤ 3

6. x = 1, y = 5 sin t, −π/2 ≤ t ≤ π/2

7. x = t2
, y = 2, −2 ≤ t ≤ 2

8. x = t3
− t + 3, y =t
2
+ 1, −2 ≤ t ≤ 2

In Exercises 9-17, sketch the graph of the given parametric equation; using a graphing utility is advisable. Be sure to
indicate the orientation of the graph.
9. x = t3
− 2t ,
2
y =t ,
2
−2 ≤ t ≤ 3

10. x = 1/t, y = sin t, 0 ≤ t ≤ 10

11. x = 3 cos t, y = 5 sin t, 0 ≤ t ≤ 2π

12. x = 3 cos t, y = 5 sin t + 3, 0 ≤ t ≤ 2π

13. x = cos t, y = cos(2t), 0 ≤t ≤π

14. x = cos t, y = sin(2t), 0 ≤ t ≤ 2π

15. x = 2 sec t, y = 3 tan t, −π/2 ≤ t ≤ π/2

16. x = cos t + 1

4
cos(8t), y = sin t +
1

4
sin(8t), 0 ≤ t ≤ 2π

17. x = cos t + 1

4
sin(8t), y = sin t +
1

4
cos(8t), 0 ≤ t ≤ 2π

In Exercises 18-19, four sets of parametric equations are given. Describe how their graphs are similar and different. Be sure
to discuss orientation and ranges.
18.
(a) x = t y = t , −∞ < t < ∞
2

(b) x = sin t y = sin t, −∞ < t < ∞


2

(c) x = e y = e , −∞ < t < ∞


t 2t

(d) x = −t y = t , −∞ < t < ∞


2

19.
(a) x = cos t y = sin t, 0 ≤ t ≤ 2π
(b) x = cos(t ) y = sin(t ), 0 ≤ t ≤ 2π
2 2

(c) x = cos(1/t) y = sin(1/t), 0 ≤ t ≤ 2π


(d) x = cos(cos t) y = sin(cos t), 0 ≤ t ≤ 2π
In Exercises 20-29, eliminate the parameter in the given parametric equations.

9.E.4 https://math.libretexts.org/@go/page/4771
20. x = 2t + 5, y = −3t + 1

21. x = sec t, y = tan t

22. x = 4 sin t + 1, y = 3 cos t − 2

23. x = t 2
, y =t
3

24. x = 1

t+1
, y =
3t+5

t+1

25. x = e t
, y =e
3t
−3

26. x = ln t, y =t
2
−1

27. x = cot t, y = csc t

28. x = cosh t, y = sinh t

29. x = cos(2t), y = sin t

In Exercises 30-33, eliminate the parameter in the given parametric equations. Describe the curve defined by the
parametric equations based on its rectangular form.
30. x = at + x 0, y = bt + y0

31. x = r cos t, y = r sin t

32. x = a cos t + h, y = b sin t + k

33. x = a sec t + h, y = b tan t + k

dy
In Exercises 34-37, find parametric equations for the given rectangular equation using the parameter t = dx
. Verify that at
t = 1 , the point on the graph has a tangent line with slope of 1.

34. y = 3x 2
− 11x + 2

35. y = e x

36. y = sin x on [0, π]


37. y = √−
x on [0, π]

In Exercises 42-45, find the value(s) of t where the graph of the parametric equations crosses itself.
38. x = t 3
− t + 3, y =t
2
−3

39. x = t 3
− 4t
2
+ t + 7, y =t
2
−t

40. x = cos t, y = sin(2t) on [0, 2π]

41. x = cos t cos(3t), y = sin t cos(3t) on [0, π]

In Exercises 42-45, find the value(s) of t where the curve defined by the parametric equations is not smooth.
42. x = t 3
+t
2
− t, y =t
2
+ 2t + 3

43. x = t 2
− 4t, y =t
3
− 2t
2
− 4t

44. x = cos t, y = 2 cos t

45. x = 2 cos t − cos(2t), y = 2 sin t − sin(2t)

In Exercises 46-54, find parametric equations that describe the given situation.
46. A projectile is fired from a height of 0ft, landing 16ft away in 4s.
47. A projectile is fired from a height of 0ft, landing 200ft away in 4s.
48. A projectile is fired from a height of 0ft, landing 200ft away in 20s.
49. A circle of radius 2, centered at the origin, that is traced clockwise once on [0, 2π].
50. A circle of radius 3, centered at (1, 1), that is traced once counter-clockwise once on [0, 1].

9.E.5 https://math.libretexts.org/@go/page/4771
51. An ellipse centered at (1,3) with vertical major axis of length 6 and minor axis of length 2.
52. An ellipse with foci at (±1, 0) and vertices at (±5, 0).
53. A hyperbola with foci at (5, −3) and (−1, −3), and with vertices at (1, −3) and (3, −3).
54. A hyperbola with vertices at (0, ±6) and asymptotes y = ±3x .

9.3: Calculus and Parametric Equations


Terms and Concepts
dy
1. T/F: Given parametric equations x = f (t) and y = g(t) , dx

= f (t)f (t)

, as long as g ′
(t) ≠ 0 .
dy
2. Given parametric equations x = f (t) and y = g(t) , the derivative dx
as given in Key Idea 37 is a function of _________?
2
d y dy
3. T/F: Given parametric equations x = f (t) and y = g(t) , to find dx
2
, one simply computes d

dt
(
dx
) .
dy
4. T/F: If dx
=0 at t = t , then the normal line to the curve at t = t is a vertical line.
0 0

Problems
In Exercises 5-12, parametric equations for a curve are given.
dy
(a) Find . dx

(b) Find the equations of the tangent and normal line(s) at the point(s) given.
(c) Sketch the graph of the parametric functions along with the found tangent and normal lines.
5. x = t, y =t ;
2
t =1

6. x = √t, y = 5t + 2; t =4

7. x = t 2
− t, y = t
2
+ t; t =1

8. x = t 2
, y =t
3
− 1; t = 0 and t = 1

9. x = sec t, y = tan t on (−π/2, π/2); t = π/4

10. x = cos t, y = sin(2t); t = π/4

11. x = cos t sin(2t), y = sin t sin(2t) on [0, 2π]; t = 3π/4

12. x = e t/10
, y =e
t/10
sin t; t = π/2

In Exercises 13-20, find t-values where the curve defined by the given parametric equations has a horizontal tangent line.
Note: these are the same equations as in Exercises 5-12.
13. x = t, y =t
2

14. x = √t, y = 5t + 2

15. x = t 2
− t, y = t
2
+t

16. x = t 2
− 1, y = t
3
−t

17. x = sec t, y = tan t on (−π/2, π/2)

18. x = cos t, y = sin(2t) on [0, 2π]

19. x = cos t sin(2t), y sin t sin(2t) on [0, 2π]

20. x = e t/10
cos t, y = e
t/10
sin t

dy
In Exercises 21-24, find t = t where the graph of the given parametric equations is not smooth, then find
0 lim
dx
.
t→t0

21. x = 2
t +1
1
, y =t
3

22. x = −t 3
+ 7t
2
− 16t + 13, y =t
3
− 5t
2
+ 8t − 2

23. x = t 3
− 3t
2
+ 3t − 1, y =t
2
− 2t + 1

9.E.6 https://math.libretexts.org/@go/page/4771
24. x = cos 2
t, y = 1 − sin
2
t

2
d y
In Exercises 25-32, parametric equations for a curve are given. Find , then determine the intervals on which the graph
dx
2

of the curve is concave up/down. Note: these are the same equations as in Exercises 5-12.
25. x = t, y =t ;
2
t =1

26. x = √t, y = 5t + 2; t =4

27. x = t2
− t, y = t
2
+ t; t =1

28. x = t2
, y =t
3
− 1; t = 0 and t = 1

29. x = sec t, y = tan t on (−π/2, π/2); t = π/4

30. x = cos t, y = sin(2t); t = π/4

31. x = cos t sin(2t), y = sin t sin(2t) on [0, 2π]; t = 3π/4

32. x = e t/10
, y =e
t/10
sin t; t = π/2

In Exercises 37-40, numerically approximates the given arc length.


37. Approximate the arc length of one petal of the rose curve x = cos t cos(2t) using Simpson's Rule and n = 4 .
38. Approximate the arc length of the "bow tie curve" x = cos t, y = sin(2t) using Simpson's Rule and n = 6 .
39. Approximate the arc length of the parabola x = t 2
−t ,y =t 2
+t on [-1,1] using Simpson's Rule and n = 4 .
−−−−
2
a2 +b
40. A common approximate of the circumference of an ellipse given by x = a cos t, y = b sin t is C ≈ 2π √
2
.

In Exercises 41-44, a solid of revolution is described. Find or approximate its surface area as specified.
41. Find the surface area of the sphere formed by rotating the circle x = 2 cos t, y = 2 sin t about:
(a) the x-axis and
(b) the y-axis.
42. Find the surface are of the torus (or "donut") formed by rotating the circle x = cos t + 2, y = sin t about the y-axis.
43. Approximate the surface are of the solid formed by rotating the "upper right half" of the bow tie curve
x = cos t, y = sin(2t) on [0, π/2] about the x-axis, using Simpson's Rule and n = 4 .

44. Approximate the surface area of the solid formed by rotating the one petal of the rose curve
x = cos t cos(2t), y = sin t cos(2t) on [0, π/4] about the x-axis, using Simpson's Rule and n = 4 .

9.4: Introduction to Polar Coordinates


Terms and Concepts
1. In your own words, describe how to plot the polar point P (r, θ).
2. T/F: When plotting a point with polar coordinateP (r, θ), r must be positive.
3. T/F: Every point in the Cartesian plane can be represented by a polar coordinate.
4. T/F: Every point in the Cartesian plane can be represented uniquely by a polar coordinate.

Problems
5. Plot the points with the given polar coordinates.
(a) A = P (2, 0)
(b) B = P (1, π)
(c) C = P (−2, π/2)
(d) D = P (1, π/4)
6. Plot the points with the given polar coordinates.
(a) A = P (2, 3π)
(b) B = P (1, −π)

9.E.7 https://math.libretexts.org/@go/page/4771
(c) C = P (1, 2)
(d) D = P (1/2, 5π/6)
7. For each of the given points give two sets of polar coordinates that identify it, where 0 ≤ θ ≤ 2π .

8. For each of the given points give two sets of polar coordinates that identify it, where −π ≤ θ ≤ π .

9. Convert each of the following polar coordinates to rectangular, and each of the following rectangular coordinates to polar.
(a) A = P (2, π/4)
(b) B = P (2, −π/4)
(c) C = P (2, −1)
(d) D = P (−2, 1)
10. Convert each of the following polar coordinates to rectangular, and each of the following rectangular coordinates to polar.
(a) A = P (3, π)
(b) B = P (1, 2π/3)
(c) C = P (0, 4)

(d) D = P (1, −√3)
In Exercises 11-29, graph the polar function on the given interval.
11. r = 2, 0 ≤ θ ≤ π/2

12. θ = π/6, −1 ≤ r ≤ π/2

13. r = 1 − cos θ, [0, 2π]

14. r = 2 + sin θ, [0, 2π]

15. r = 2 − sin θ, [0, 2π]

16. r = 1 − 2 sin θ, [0, 2π]

17. r = 1 + 2 sin θ, [0, 2π]

18. r = cos(2θ), [0, 2π]

19. r = sin(3θ), [0, π]

20. r = cos(θ/3), [0, 3π]

21. r = cos(2θ/3), [0, 6π]

22. r = θ/2, [0, 4π]

23. r = 3 sin(θ), [0, π]

24. r = cos θ sin θ, [0, 2π]

25. r = θ 2
− (π/2 ) ,
2
[π, π]

26. r = 3

5 sin θ−cos θ
, [0, 2π]

27. r = −2

3 cos θ−2 sin θ


, [0, 2π]

9.E.8 https://math.libretexts.org/@go/page/4771
28. r = 3 sec θ, (−π/2, π/2)

29. r = 3 csc θ, (0, π)

In Exercises 30-38, convert the polar equation to a rectangular equation.


30. r = 2 cos θ
31. r = −4 sin θ
32. r = cos θ + sin θ
33. r = 7

5 sin θ−2 cos θ

34. r = 3

cos θ

35. r = 4

sin θ

36. r = tan θ
37. r = 2
38. θ = π/6
In Exercises 39-46, convert the rectangular equation to a polar equation.
39. y = x
40. y = 4x + 7
41. x = 5
42. y = 5
43. x = y 2

44. x2
y =1

45. x2
+y
2
=7

46. x + 1) 2
+y
2
=1

In Exercises 47-54, find the points of intersection of the polar graphs.


47. r = sin(2θ) and r = cos θ on [0, π] .
48. r = cos(2θ) and r = cos θ on [0, π] .
49. r = 2 cos θ and r = 2 sin θ on [0, π] .

50. r = sin θ and r = √3 + 3 sin θ on [0, 2π] .
51. r = sin(3θ) and r = cos(3θ) on [0, π] .
52. r = 3 cos θ and r = 1 + cos θ on [−π, π] .
53. r = 1 and r = 2 sin(2θ) on [0, 2π] .
54. r = 1 − cos θ and r = 1 + sin θ on [0, 2π] .
55. Pick an integer value for n, where n ≠ 2, 3 and use technology to plot r = sin( m

n
θ) for three different integer values of m.
Sketch these and determine a minimal interval on which the entire graph is shown.
56. Create your own polar function, r = f (θ) and sketch it. Describe why the graph looks as it does.

9.5: Calculus and Polar Functions


Terms and Concepts
1. Given polar equation r = f (θ) , how can one create parametric equations of the same curve?
2. With rectangular coordinates, it is natural to approximate are with ________; with polar coordinates, it is natural to approximate
area with _______.

9.E.9 https://math.libretexts.org/@go/page/4771
Problems
In Exercises 3-10, find;
dy
(a) dx

(b) the equation of the tangent and normal lines to the curve at the indicated θ -value.
3. r = 1; θ = π/4

4. r = cos θ; θ = π/4

5. r = 1 + sin θ; θ = π/6

6. r = 1 − cos θ; θ = 3π/4

7. r = θ; θ = π/2

8. r = cos(3θ); θ = π/6

9. r = sin(4θ); θ = π/3

10. r = 1

sin θ−cos θ
; θ =π

In Exercises 11-14, find the values of θ in the given interval where the graph of the polar function has horizontal and
vertical tangent lines.
11. r = 3; [0, 2π]

12. r = 2 sin θ; [0, π]

13. r = cos(2θ); [0, 2π]

14. r = 1 + cos θ; [0, 2π]

In Exercises 15-16, find the equation of the lines tangent to the graph at the pole.
15. r = sin θ; [0, π]

16. r = sin(3θ); [0, π]

In Exercises 17-27, find the area of the described region.


17. Enclosed by the circle: r = 4 sin θ
18. Enclosed by the circle r = 5
19. Enclosed by one petal of r = sin(3θ)
20. Enclosed by the cardiod r = 1 − sin θ
21. Enclosed by the inner loop of the limacon r = 1 + 2 cos t
22. Enclosed by the outer loop of the limacon r = 1 + 2 cos t (including area enclosed by the inner loop)
23. Enclosed between the inner and outer loop of the limacon r = 1 + 2 cos t
24. Enclosed by r = 2 cos θ and r = sin θ , as shown:

9.E.10 https://math.libretexts.org/@go/page/4771
25. Enclosed by r = cos(3θ) and r = sin(3θ) , as shown:

26. Enclosed by r = cos θ and r = sin(2θ) , as shown:

27. Enclosed by r = cos θ and r = 1 − cos θ , as shown:

In Exercises 28-32, answer the questions involving arc length.


28. Let x(θ) = f (θ) cos θ and y(θ) = f (θ) sin θ . Show, as suggested by the text, that x (θ) ′ 2 ′
+ y (θ)
2 ′
= f (θ)
2 2
+ f (θ) .
29. Use the arc length formula to compute the arc length of the circle r = 2 .
30. Use the arc length formula to compute the arc length of the circle r = 4 sin θ .
31. Approximate the arc length of one petal of the rose curve r = sin(3θ) with Simpson's Rule and n = 4 .
32. Approximate the arc length of the cardiod r = 1 + cos θ with Simpson's Rule and n = 6 .
In Exercises 33-37, answer the questions involving surface area.
33. Use Key Idea 44 to find the surface area of the sphere formed by revolving the circle r = 2 about the initial ray.
34. Use Key Idea 44 to find the surface area of the sphere formed by revolving the circle r = 2 cos θ about the initial ray.
35. Find the surface area of a solid formed by revolving the cardiod r = 1 + cos θ about the initial ray.
36. Find the surface area of the solid formed by revolving the circle r = 2 cos θ about the line θ = π/2 .
37. Find the surface area of the solid formed by revolving the line r = 3 sec θ, −π/4 ≤ θ ≤ π/4 , about the line θ = π/2 .

This page titled 9.E: Applications of Curves in a Plane (Exercises) is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or
curated by Gregory Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history
is available upon request.

9.E.11 https://math.libretexts.org/@go/page/4771
CHAPTER OVERVIEW
10: Vectors
This chapter introduces a new mathematical object, the vector. We will see that vectors provide a powerful language for describing
quantities that have magnitude and direction aspects. A simple example of such a quantity is force: when applying a force, one is
generally interested in how much force is applied (i.e., the magnitude of the force) and the direction in which the force was applied.
Vectors will play an important role in many of the subsequent chapters in this text. This chapter begins with moving our
mathematics out of the plane and into "space.'' That is, we begin to think mathematically not only in two dimensions, but in three.
With this foundation, we can explore vectors both in the plane and in space.
10.1: Introduction to Cartesian Coordinates in Space
10.2: An Introduction to Vectors
10.3: The Dot Product
10.4: The Cross Product
10.5: Lines
10.6: Planes
10.E: Applications of Vectors (Exercises)

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 10: Vectors is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et al. via
source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

1
10.1: Introduction to Cartesian Coordinates in Space
Up to this point in this text we have considered mathematics in a 2--dimensional world. We have plotted graphs on the x-y plane
using rectangular and polar coordinates and found the area of regions in the plane. We have considered properties of solid objects,
such as volume and surface area, but only by first defining a curve in the plane and then rotating it out of the plane.
While there is wonderful mathematics to explore in "2D,'' we live in a "3D'' world and eventually we will want to apply
mathematics involving this third dimension. In this section we introduce Cartesian coordinates in space and explore basic surfaces.
This will lay a foundation for much of what we do in the remainder of the text.
Each point P in space can be represented with an ordered triple, P = (a, b, c) , where a , b and c represent the relative position of
P along the x-, y - and z -axes, respectively. Each axis is perpendicular to the other two.

Visualizing points in space on paper can be problematic, as we are trying to represent a 3-dimensional concept on a 2--dimensional
medium. We cannot draw three lines representing the three axes in which each line is perpendicular to the other two. Despite this
issue, standard conventions exist for plotting shapes in space that we will discuss that are more than adequate.
One convention is that the axes must conform to the right hand rule. This rule states that when the index finger of the right hand is
extended in the direction of the positive x-axis, and the middle finger (bent "inward'' so it is perpendicular to the palm) points along
the positive y -axis, then the extended thumb will point in the direction of the positive z -axis. (It may take some thought to verify
this, but this system is inherently different from the one created by using the "left hand rule.''). There are two popular methods to
draw axes that we briefly discuss.
In Figure 10.1.1 we see the point P = (2, 1, 3) plotted on a set of axes. The basic convention here is that the x-y plane is drawn in
its standard way, with the z -axis down to the left. The perspective is that the paper represents the x-y plane and the positive z axis
is coming up, off the page. This method is preferred by many engineers. Because it can be hard to tell where a single point lies in
relation to all the axes, dashed lines have been added to let one see how far along each axis the point lies.

Figure 10.1.1 : Plotting the point P = (2, 1, 3) in space.


One can also consider the x-y plane as being a horizontal plane in, say, a room, where the positive z -axis is pointing up. When one
steps back and looks at this room, one might draw the axes as shown in Figure 10.1.2. The same point P is drawn, again with
dashed lines. This point of view is preferred by most mathematicians, and is the convention adopted by this text.

10.1.1 https://math.libretexts.org/@go/page/4213
Figure 10.1.2 : Plotting the point P = (2, 1, 3) in space with a perspective used in this text.

Note
As long as the coordinate axes are positioned so that they follow the right hand rule, it does not matter how the axes are drawn
on paper.

Measuring Distances
It is of critical importance to know how to measure distances between points in space. The formula for doing so is based on
measuring distance in the plane, and is known (in both contexts) as the Euclidean measure of distance.

Definition 48: distance in space

Let P = (x1 , y1 , z1 ) and Q = (x 2, y2 , z2 ) be points in space. The distance D between P and Q is


−−−−−−−−−−−−−−−−−−−−−−−−−−−−
2 2 2
D = √ (x2 − x1 ) + (y2 − y1 ) + (z2 − z1 ) . (10.1.1)

¯
¯¯¯¯¯¯
¯ ¯
¯¯¯¯¯¯
¯
We refer to the line segment that connects points P and Q in space as P Q, and refer to the length of this segment as || P Q|| . The
above distance formula allows us to compute the length of this segment.

Example 10.1.1: Length of a line segment


¯
¯¯¯¯¯¯
¯
Let P = (1, 4, −1) and let Q = (2, 1, 1). Draw the line segment P Q and find its length.
Solution
The points P and Q are plotted in Figure 10.1.3; no special consideration need be made to draw the line segment connecting
these two points; simply connect them with a straight line.

Figure 10.1.3 : Plotting points P and Q in Example 10.1.1 .


One cannot actually measure this line on the page and deduce anything meaningful; its true length must be measured
analytically. Applying Definition 48, we have

10.1.2 https://math.libretexts.org/@go/page/4213
−−−−−−−−−−−−−−−−−−−−−−−−−−
¯
¯¯¯¯¯¯
¯ 2 2 2 −−
|| P Q|| = √ (2 − 1 ) + (1 − 4 ) + (1 − (−1)) = √14 ≈ 3.74. (10.1.2)

Spheres
Just as a circle is the set of all points in the plane equidistant from a given point (its center), a sphere is the set of all points in space
that are equidistant from a given point. Definition 48 allows us to write an equation of the sphere. We start with a point
C = (a, b, c) which is to be the center of a sphere with radius r . If a point P = (x, y, z) lies on the sphere, then P is r units from

C ; that is,

−−−−−−−−−−−−−−−−−−−−−−−
¯
¯¯¯¯¯¯
¯ 2 2 2
|| P C || = √ (x − a) + (y − b ) + (z − c ) = r. (10.1.3)

Squaring both sides, we get the standard equation of a sphere in space with center at C = (a, b, c) with radius r, as given in the
following Key Idea.

KEY IDEA 45: STANDARD EQUATION OF A SPHERE IN SPACE


The standard equation of the sphere with radius r, centered at C = (a, b, c) , is
2 2 2 2
(x − a) + (y − b ) + (z − c ) =r . (10.1.4)

Example 10.1.2: Equation of a sphere

Find the center and radius of the sphere defined by x 2


+ 2x + y
2
− 4y + z
2
− 6z = 2.

Solution
To determine the center and radius, we must put the equation in standard form. This requires us to complete the square (three
times).
2 2 2
x + 2x + y − 4y + z − 6z =2
2 2 2
(x + 2x + 1) + (y − 4y + 4) + (z − 6z + 9) − 14 =2

2 2 2
(x + 1 ) + (y − 2 ) + (z − 3 ) = 16

The sphere is centered at (−1, 2, 3) and has a radius of 4.

The equation of a sphere is an example of an implicit function defining a surface in space. In the case of a sphere, the variables x, y
and z are all used. We now consider situations where surfaces are defined where one or two of these variables are absent.

Introduction to Planes in Space


The coordinate axes naturally define three planes (shown in Figure 10.1.4), the coordinate planes: the x-y plane, the y -z plane
and the x-z plane. The x-y plane is characterized as the set of all points in space where the z -value is 0. This, in fact, gives us an
equation that describes this plane: z = 0 . Likewise, the x-z plane is all points where the y -value is 0, characterized by y = 0 .

Figure 10.1.4 : The coordinate planes.


The equation x = 2 describes all points in space where the x-value is 2. This is a plane, parallel to the y -z coordinate plane, shown
in Figure 10.1.5.

10.1.3 https://math.libretexts.org/@go/page/4213
Figure 10.1.5 : The plane x = 2 .

Example 10.1.3: Regions defined by planes

Sketch the region defined by the inequalities −1 ≤ y ≤ 2 .


Solution
The region is all points between the planes y = −1 and y = 2 . These planes are sketched in Figure 10.1.6, which are parallel
to the x-z plane. Thus the region extends infinitely in the x and z directions, and is bounded by planes in the y direction.

Figure 10.1.6 : Sketching the boundaries of a region in Example 10.1.3 .

Cylinders
The equation x = 1 obviously lacks the y and z variables, meaning it defines points where the y and z coordinates can take on any
value. Now consider the equation x + y = 1 in space. In the plane, this equation describes a circle of radius 1, centered at the
2 2

origin. In space, the z coordinate is not specified, meaning it can take on any value. In Figure 10.1.7a, we show part of the graph
of the equation x + y = 1 by sketching 3 circles: the bottom one has a constant z -value of −1.5, the middle one has a z -value of
2 2

0 and the top circle has a z -value of 1. By plotting all possible z -values, we get the surface shown in Figure 10.1.7b.
This surface looks like a "tube,'' or a "cylinder''; mathematicians call this surface a cylinder for an entirely different reason.

10.1.4 https://math.libretexts.org/@go/page/4213
Figure 10.1.7 : Sketching x 2
+y
2
= 1 .

Definition 49: CYLINDER, Directrix and Rulings


Let C be a curve in a plane and let L be a line not parallel to C . A cylinder is the set of all lines parallel to L that pass through
C . The curve C is the directrix of the cylinder, and the lines are the rulings.

In this text, we consider curves C that lie in planes parallel to one of the coordinate planes, and lines L that are perpendicular to
these planes, forming right cylinders. Thus the directrix can be defined using equations involving 2 variables, and the rulings will
be parallel to the axis of the 3 variable.
rd

In the example preceding the definition, the curve x + y = 1 in the x-y plane is the directrix and the rulings are lines parallel to
2 2

the z -axis. (Any circle shown in Figure 10.8 can be considered a directrix; we simply choose the one where z = 0 .) Sample rulings
can also be viewed in part (b) of the figure. More examples will help us understand this definition.

Example 10.1.4: Graphing cylinders

Graph the cylinder following cylinders.


1. z = y 2

2. x = sin z
Solution
1. We can view the equation z = y as a parabola in the y -z plane, as illustrated in Figure 10.1.8a. As x does not appear in
2

the equation, the rulings are lines through this parabola parallel to the x-axis, shown in Figure 10.1.8b. These rulings give a
general idea as to what the surface looks like, drawn in (c).

Figure 10.1.8 : Sketching the cylinder defined by z = y


2
.
2. We can view the equation x = sin z as a sine curve that exists in the x-z plane, as shown in Figure 10.1.9a. The rules are
parallel to the y axis as the variable y does not appear in the equation x = sin z ; some of these are shown in Figure 10.1.9b.
The surface is shown in part (c) of the figure.

10.1.5 https://math.libretexts.org/@go/page/4213
Figure 10.1.9 : Sketching the cylinder defined by x = sin z .

Surfaces of Revolution
One of the applications of integration we learned previously was to find the volume of solids of revolution -- solids formed by
revolving a curve about a horizontal or vertical axis. We now consider how to find the equation of the surface of such a solid.
Consider the surface formed by revolving y = √− x about the x-axis. Cross--sections of this surface parallel to the y -z plane are

circles, as shown in Figure 10.1.1a. Each circle has equation of the form y + z = r for some radius r. The radius is a function
2 2 2

of x; in fact, it is r(x) = √− 2 −
x . Thus the equation of the surface shown in Figure 10.1.10b is y + z = (√x ) .
2 2

Figure 10.1.10: Introducing surfaces of revolution.


We generalize the above principles to give the equations of surfaces formed by revolving curves about the coordinate axes.

KEY IDEA 46: SURFACES OF REVOLUTION, PART 1


Let r be a radius function.
1. The equation of the surface formed by revolving y = r(x) or z = r(x) about the x-axis is y 2
+z
2
= r(x )
2
.
2. The equation of the surface formed by revolving x = r(y) or z = r(y) about the y -axis is x 2
+z
2
= r(y )
2
.
3. The equation of the surface formed by revolving x = r(z) or y = r(z) about the z -axis is x 2
+y
2
= r(z)
2
.

Example 10.1.5: Finding equation of a surface of revolution

Let y = sin z on [0, π]. Find the equation of the surface of revolution formed by revolving y = sin z about the z -axis.
Solution
Using Key Idea 46, we find the surface has equation x
2
+y
2 2
= sin z . The curve is sketched in Figure 10.1.11a and the
surface is drawn in Figure 10.1.11b.
Note how the surface (and hence the resulting equation) is the same if we began with the curve x = sin z , which is also drawn
in Figure 10.1.11a.

10.1.6 https://math.libretexts.org/@go/page/4213
Figure 10.1.11: Revolving y = sin z about the z-axis in Example 10.1.5 .

This particular method of creating surfaces of revolution is limited. For instance, in Section 7.3 we found the volume of the solid
formed by revolving y = sin x about the y -axis. Our current method of forming surfaces can only rotate y = sin x about the x-
axis. Trying to rewrite y = sin x as a function of y is not trivial, as simply writing x = sin y only gives part of the region we
−1

desire.
What we desire is a way of writing the surface of revolution formed by rotating y = f (x) about the y -axis. We start by first
recognizing this surface is the same as revolving z = f (x) about the z -axis. This will give us a more natural way of viewing the
surface.
A value of x is a measurement of distance from the z -axis. At the distance r, we plot a z -height of f (r). When rotating f (x) about
the z -axis, we want all points a distance of r from the z -axis in the x-y plane to have a z -height of f (r). All such points satisfy the
−−− −− − −−− −− − − −− −−−
equation r = x + y ; hence r = √x + y . Replacing r with √x + y in f (r) gives z = f (√x + y ) . This is the equation
2 2 2 2 2 2 2 2 2

of the surface.

KEY IDEA 47: SURFACES OF REVOLUTION, PART 2


Let z = f (x) , x ≥ 0 , be a curve in the x-z plane. The surface formed by revolving this curve about the z -axis has equation
−−− −− −
z = f (√x + y ) .
2 2

Example 10.1.6: Finding equation of surface of revolution

Find the equation of the surface found by revolving z = sin x about the z -axis.
Solution
−−−−−−
Using Key Idea 47, the surface has equation z = sin(√x 2 2
+y ) . The curve and surface are graphed in Figure 10.1.12.

10.1.7 https://math.libretexts.org/@go/page/4213
Figure 10.1.12: Revolving z = sin x about the z-axis in Example 10.1.6 .

Quadric Surfaces
Spheres, planes and cylinders are important surfaces to understand. We now consider one last type of surface, a quadric surface.
The definition may look intimidating, but we will show how to analyze these surfaces in an illuminating way.

Definition 50: QUADRIC SURFACE


A quadric surface is the graph of the general second--degree equation in three variables:
2 2 2
Ax + By + Cz + Dxy + Exz + F yz + Gx + H y + I z + J = 0. (10.1.5)

When the coefficients D, E or F are not zero, the basic shapes of the quadric surfaces are rotated in space. We will focus on
quadric surfaces where these coeffiecients are 0; we will not consider rotations. There are six basic quadric surfaces: the elliptic
paraboloid, elliptic cone, ellipsoid, hyperboloid of one sheet, hyperboloid of two sheets, and the hyperbolic paraboloid.
We study each shape by considering traces, that is, intersections of each surface with a plane parallel to a coordinate plane. For
instance, consider the elliptic paraboloid z = x /4 + y , shown in Figure 10.13. If we intersect this shape with the plane z = d
2 2

(i.e., replace z with d ), we have the equation:


2
x 2
d = +y . (10.1.6)
4

Divide both sides by d :


2 2
x y
1 = + . (10.1.7)
4d d

This describes an ellipse -- so cross sections parallel to the x-y coordinate plane are ellipses. This ellipse is drawn in Figure
10.1.13.

10.1.8 https://math.libretexts.org/@go/page/4213
Figure 10.1.13: The elliptic paraboloid z 2
= x /4 + y
2
.
Now consider cross sections parallel to the x-z plane. For instance, letting y = 0 gives the equation z = x /4 , clearly a parabola.2

Intersecting with the plane x = 0 gives a cross section defined by z = y , another parabola. These parabolas are also sketched in
2

the figure.
Thus we see where the elliptic paraboloid gets its name: some cross sections are ellipses, and others are parabolas.
Such an analysis can be made with each of the quadric surfaces. We give a sample equation of each, provide a sketch with
representative traces, and describe these traces.
2
2
y
Elliptic Paraboloid, z = x
2
a
+ 2
b

One variable in the equation of the elliptic paraboloid will be raised to the first power; above, this is the z variable. The paraboloid
will "open'' in the direction of this variable's axis. Thus x = y /a + z /b is an elliptic paraboloid that opens along the x-axis.
2 2 2 2

Multiplying the right hand side by (−1) defines an elliptic paraboloid that "opens'' in the opposite direction.
2
2
y
Elliptic Cone, z 2
=
x
2
a
+ 2
b

10.1.9 https://math.libretexts.org/@go/page/4213
One can rewrite the equation as z 2
− x /a
2 2 2
− y /b
2
=0 . The one variable with a positive coefficient corresponds to the axis that
the cones "open'' along.
2
2 y 2

Ellipsoid, x
2
a
+ 2
+
z

c
2
=1
b

If a = b = c ≠ 0 , the ellipsoid is a sphere with radius a ; compare to Key Idea 45.


2
2 y 2

Hyperboloid of One Sheet, x


2
a
+ 2

z

c
2
=1
b

The one variable with a negative coefficient corresponds to the axis that the hyperboloid "opens'' along.
2
2 2 y
Hyperboloid of Two Sheets, z

c
2

x

a
2
− 2
=1
b

10.1.10 https://math.libretexts.org/@go/page/4213
The one variable with a positive coefficient corresponds to the axis that the hyperboloid "opens'' along. In the case illustrated, when
|d| < |c| , there is no trace.

2
2 y
Hyperbolic Paraboloid, z = x
2
a
− 2
b

The parabolic traces will open along the axis of the one variable that is raised to the first power.

Example 10.1.7: Sketching quadric surfaces

Sketch the quadric surface defined by the given equation.


2 2

1. y = x

4
+
z

16
2
y 2

2. x +
2

9
+
z

4
= 1.

3. z = y 2
−x
2
.
Solution

10.1.11 https://math.libretexts.org/@go/page/4213
2 2

1. y = x

4
+
z

16
:

We first identify the quadric by pattern--matching with the equations given previously. Only two surfaces have equations
where one variable is raised to the first power, the elliptic paraboloid and the hyperbolic paraboloid. In the latter case, the
other variables have different signs, so we conclude that this describes a hyperbolic paraboloid. As the variable with the
first power is y , we note the paraboloid opens along the y -axis.

To make a decent sketch by hand, we need only draw a few traces. In this case, the traces x = 0 and z = 0 form parabolas
that outline the shape.

x =0 : The trace is the parabola y = z 2


/16

z =0 : The trace is the parabola y = x 2


/4 .

Graphing each trace in the respective plane creates a sketch as shown in Figure 10.1.14a. This is enough to give an idea of
what the paraboloid looks like. The surface is filled in in Figure 10.1.14b.

Figure 10.1.14 : Sketching an elliptic paraboloid.


2
2
y
2. x
2
+
9
+
z

4
=1 :

This is an ellipsoid. We can get a good idea of its shape by drawing the traces in the coordinate planes.

2
2
y
x =0 : The trace is the ellipse + = 1 . The major axis is along the y --axis with length 6 (as b = 3 , the length of the
9
z

axis is 6); the minor axis is along the z -axis with length 4.

y =0 : The trace is the ellipse x 2


+
z

4
= 1. The major axis is along the z -axis, and the minor axis has length 2 along the
x-axis.

2
y
z =0 : The trace is the ellipse x 2
+
9
= 1, with major axis along the y -axis.

Graphing each trace in the respective plane creates a sketch as shown in Figure Figure 10.1.15a. Filling in the surface gives
Figure Figure 10.1.15b.

10.1.12 https://math.libretexts.org/@go/page/4213
Figure 10.1.15 : Sketching an ellipsoid.
3. z = y − x :
2 2

This defines a hyperbolic paraboloid, very similar to the one shown in the gallery of quadric sections. Consider the traces in
the y − z and x − z planes:

x =0 : The trace is z = y , a parabola opening up in the y − z plane.


2

y =0 : The trace is z = −x , a parabola opening down in the x − z plane.


2

Sketching these two parabolas gives a sketch like that in Figure Figure 10.1.16a, and filling in the surface gives a sketch
like Figure 10.1.16b.

Figure 10.1.16: Sketching a hyperbolic paraboloid.

10.1.13 https://math.libretexts.org/@go/page/4213
Example 10.1.8: Identifying quadric surfaces

Consider the quadric surface shown in Figure Figure 10.1.17. Which of the following equations best fits this surface?

Figure 10.1.8 .
2
2 2
z 2 2 2
(a) x −y − =0 (c) z −x −y =1
9
2
z
2 2 2 2 2
(b) x −y −z =1 (d) 4 x −y − =1
9

Solution
The image clearly displays a hyperboloid of two sheets. The gallery informs us that the equation will have a form similar to
2
2 2 y
z

c
2

x
2
a
− 2
=1 .
b

We can immediately eliminate option (a), as the constant in that equation is not 1.
The hyperboloid "opens'' along the x-axis, meaning x must be the only variable with a positive coefficient, eliminating (c).
The hyperboloid is wider in the z -direction than in the y -direction, so we need an equation where c >b . This eliminates (b),
2

leaving us with (d). We should verify that the equation given in (d), 4x − y − = 1 , fits. 2 2 z

We already established that this equation describes a hyperboloid of two sheets that opens in the x-direction and is wider in the
2

z -direction than in the y . Now note the coefficient of the x-term. Rewriting 4x in standard form, we have: 4 x = . Thus 2 2 x
2
(1/2)

when y = 0 and z = 0 , x must be 1/2; i.e., each hyperboloid "starts'' at x = 1/2. This matches our figure.
2

We conclude that 4x 2
−y
2

z

9
=1 best fits the graph.

This section has introduced points in space and shown how equations can describe surfaces. The next sections explore vectors, an
important mathematical object that we'll use to explore curves in space.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 10.1: Introduction to Cartesian Coordinates in Space is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or
curated by Gregory Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history
is available upon request.

10.1.14 https://math.libretexts.org/@go/page/4213
10.2: An Introduction to Vectors
Many quantities we think about daily can be described by a single number: temperature, speed, cost, weight and height. There are
also many other concepts we encounter daily that cannot be described with just one number. For instance, a weather forecaster
often describes wind with its speed and its direction ("… with winds from the southeast gusting up to 30 mph …"). When
applying a force, we are concerned with both the magnitude and direction of that force. In both of these examples, direction is
important. Because of this, we study vectors, mathematical objects that convey both magnitude and direction information.
One "bare--bones'' definition of a vector is based on what we wrote above: "a vector is a mathematical object with magnitude and
direction parameters.'' This definition leaves much to be desired, as it gives no indication as to how such an object is to be used.
Several other definitions exist; we choose here a definition rooted in a geometric visualization of vectors. It is very simplistic but
readily permits further investigation.

Definition 51 Vector
A vector is a directed line segment.

Given points P and Q (either in the plane or in space), we denote with P Q the vector from P to Q. The point P is said to be
the initial point of the vector, and the point Q is the terminal point.

¯
¯¯¯¯¯¯
¯
The magnitude, length or norm of P Q is the length of the line segment P Q:

¯
¯¯¯¯¯¯
¯
∥ P Q∥ = ∥ P Q∥. (10.2.1)

Two vectors are equal if they have the same magnitude and direction.

Figure 10.18 shows multiple instances of the same vector. Each directed line segment has the same direction and length
(magnitude), hence each is the same vector.

Figure 10.18: Drawing the same vector with different initial points.
We use R (pronounced "r two'') to represent all the vectors in the plane, and use
2
R
3
(pronounced "r three'') to represent all the
vectors in space.

Figure 10.19: Illustrating how equal vectors have the same displacement.
→ →
Consider the vectors P Q and RS as shown in Figure 10.19. The vectors look to be equal; that is, they seem to have the same
length and direction. Indeed, they are. Both vectors move 2 units to the right and 1 unit up from the initial point to reach the
terminal point. One can analyze this movement to measure the magnitude of the vector, and the movement itself gives direction
information (one could also measure the slope of the line passing through P and Q or R and S ). Since they have the same length
and direction, these two vectors are equal.

10.2.1 https://math.libretexts.org/@go/page/4214
This demonstrates that inherently all we care about is displacement; that is, how far in the x, y and possibly z directions the
→ →
terminal point is from the initial point. Both the vectors P Q and RS in Figure 10.19 have an x-displacement of 2 and a y -
displacement of 1. This suggests a standard way of describing vectors in the plane. A vector whose x-displacement is a and whose
y -displacement is b will have terminal point (a, b) when the initial point is the origin, (0, 0). This leads us to a definition of a

standard and concise way of referring to vectors.

Definition 52 Component Form of a Vector


1. The component form of a vector v ⃗ in R , whose terminal point is (a,
2
b) when its initial point is (0, 0), is ⟨a, b⟩.
2. The component form of a vector v ⃗ in R , whose terminal point is (a,
3
b, c) when its initial point is (0, 0, 0), is ⟨a, b, c⟩.
The numbers a , b (and c , respectively) are the components of v⃗.


It follows from the definition that the component form of the vector P Q, where P = (x1 , y1 ) and Q = (x 2, y2 ) is

P Q = ⟨x2 − x1 , y2 − y1 ⟩; (10.2.2)


in space, where P = (x1 , y1 , z1 ) and Q = (x
2, y2 , z2 ) , the component form of P Q is

P Q = ⟨x2 − x1 , y2 − y1 , z2 − z1 ⟩. (10.2.3)

We practice using this notation in the following example.

Example 10.2.1: Using component form notation for vectors

1. Sketch the vector v ⃗ = ⟨2, −1⟩ starting at P = (3, 2) and find its magnitude.
2. Find the component form of the vector w⃗ whose initial point is R = (−3, −2) and whose terminal point is S = (−1, 2) .
3. Sketch the vector u⃗ = ⟨2, −1, 3⟩ starting at the point Q = (1, 1, 1) and find its magnitude.
Solution
1. Using P as the initial point, we move 2 units in the positive x-direction and −1 units in the positive y -direction to arrive at
the terminal point P = (5, 1) , as drawn in Figure 10.20(a).

The magnitude of v ⃗ is determined directly from the component form:


−−−−−−−−−
2 2 –
⃗ = √2
∥ v∥ + (−1 ) = √5. (10.2.4)

2. Using the note following Definition 52, we have



RS = ⟨−1 − (−3), 2 − (−2)⟩ = ⟨2, 4⟩. (10.2.5)


One can readily see from Figure 10.20(a) that the x- and y -displacement of RS is 2 and 4, respectively, as the component
form suggests.
3. Using Q as the initial point, we move 2 units in the positive x-direction, −1 unit in the positive y -direction, and 3 units in
the positive z -direction to arrive at the terminal point Q = (3, 0, 4), illustrated in Figure 10.20(b).

The magnitude of u⃗ is:


−−−−−−−−−−−−−
2 2 2 −−
∥ u⃗∥ = √ 2 + (−1 ) +3 = √14. (10.2.6)

10.2.2 https://math.libretexts.org/@go/page/4214
Figure 10.20: Graphing vectors in Example 10.2.1

Now that we have defined vectors, and have created a nice notation by which to describe them, we start considering how vectors
interact with each other. That is, we define an algebra on vectors.

Definition 53 VECTOR ALGEBRA


1. Let u⃗ = ⟨u , u ⟩ and v ⃗ = ⟨v , v ⟩ be vectors in R , and let c be a scalar.
1 2 1 2
2

(a) The addition, or sum, of the vectors u⃗ and v ⃗ is the vector


u⃗ + v ⃗ = ⟨u1 + v1 , u2 + v2 ⟩. (10.2.7)

(b) The scalar product of c and v ⃗ is the vector


c v ⃗ = c⟨v1 , v2 ⟩ = ⟨c v1 , c v2 ⟩. (10.2.8)

2. Let u⃗ = ⟨u , u , u ⟩ and v ⃗ = ⟨v , v , v ⟩ be vectors in R , and let c be a scalar.


1 2 3 1 2 3
3

(a) The addition, or sum, of the vectors u⃗ and v ⃗ is the vector


u⃗ + v⃗ = ⟨u1 + v1 , u2 + v2 , u3 + v3 ⟩. (10.2.9)

(b) The scalar product of c and v ⃗ is the vector


c v ⃗ = c⟨v1 , v2 , v3 ⟩ = ⟨c v1 , c v2 , c v3 ⟩. (10.2.10)

In short, we say addition and scalar multiplication are computed "component--wise.''

Example 10.2.2: Adding vectors

Sketch the vectors u⃗ = ⟨1, 3⟩ , v ⃗ = ⟨2, 1⟩ and u⃗ + v ⃗ all with initial point at the origin.

Solution
We first compute u⃗ + v ⃗ .
u⃗ + v⃗ = ⟨1, 3⟩ + ⟨2, 1⟩

= ⟨3, 4⟩.

10.2.3 https://math.libretexts.org/@go/page/4214
These are all sketched in Figure 10.21.

Figure 10.21: Graphing the sum of vectors in Example 10.2.2

As vectors convey magnitude and direction information, the sum of vectors also convey length and magnitude information. Adding
u⃗ + v ⃗ suggests the following idea:

"Starting at an initial point, go out u⃗, then go out v."


⃗ (10.2.11)

This idea is sketched in Figure 10.22, where the initial point of v ⃗ is the terminal point of u⃗. This is known as the "Head to Tail
Rule'' of adding vectors. Vector addition is very important. For instance, if the vectors u⃗ and v ⃗ represent forces acting on a body,
the sum u⃗ + v ⃗ gives the resulting force. Because of various physical applications of vector addition, the sum u⃗ + v ⃗ is often referred
to as the resultant vector, or just the "resultant.''

Figure 10.22: Illustrating how to add vectors using the Head to Tail Rule and Parallelogram Law.
Analytically, it is easy to see that u⃗ + v⃗ = v⃗ + u⃗ . Figure 10.22 also gives a graphical representation of this, using gray vectors.
Note that the vectors u⃗ and v⃗, when arranged as in the figure, form a parallelogram. Because of this, the Head to Tail Rule is also
known as the Parallelogram Law: the vector u⃗ + v⃗ is defined by forming the parallelogram defined by the vectors u⃗ and v⃗; the
initial point of u⃗ + v⃗ is the common initial point of parallelogram, and the terminal point of the sum is the common terminal point
of the parallelogram.
While not illustrated here, the Head to Tail Rule and Parallelogram Law hold for vectors in R as well.
3

It follows from the properties of the real numbers and Definition 53 that
u⃗ − v ⃗ = u⃗ + (−1)v.⃗ (10.2.12)

The Parallelogram Law gives us a good way to visualize this subtraction. We demonstrate this in the following example.

Example 10.2.3: Vector Subtraction

Let u⃗ = ⟨3, 1⟩ and v ⃗ = ⟨1, 2⟩. Compute and sketch u⃗ − v ⃗ .


Solution
The computation of u⃗ − v ⃗ is straightforward, and we show all steps below. Usually the formal step of multiplying by (−1) is
omitted and we "just subtract.''
u⃗ − v ⃗ = u⃗ + (−1)v ⃗

= ⟨3, 1⟩ + ⟨−1, −2⟩

= ⟨2, −1⟩.

10.2.4 https://math.libretexts.org/@go/page/4214
Figure 10.23: Illustrating how to subtract vectors graphically.

Figure 10.23 illustrates, using the Head to Tail Rule, how the subtraction can be viewed as the sum u⃗ + (−v)⃗ . The figure also
illustrates how u⃗ − v ⃗ can be obtained by looking only at the terminal points of u⃗ and v ⃗ (when their initial points are the same).

Example 10.2.4: Scaling vectors

1. Sketch the vectors v ⃗ = ⟨2, 1⟩ and 2v ⃗ with initial point at the origin.
2. Compute the magnitudes of v ⃗ and 2v.⃗
Solution
1. We compute 2v⃗:

2v ⃗ = 2⟨2, 1⟩

= ⟨4, 2⟩.

Both v ⃗ and 2v ⃗ are sketched in Figure10.24. Make note that 2v ⃗ does not start at the terminal point of v ⃗; rather, its initial
point is also the origin.
2. The figure suggests that 2v ⃗ is twice as long as v ⃗. We compute their magnitudes to confirm this.
−−−−−−
2 2

∥ v∥ = √2 + 1

= √5.
− −−−−−
2 2
∥2 v⃗∥ = √ 4 + 2
−−
= √20
−− − –
= √ 4 ⋅ 5 = 2 √5.

As we suspected, 2v⃗ is twice as long as v⃗.

Figure 10.24: Graphing vectors v ⃗ and 2v ⃗ in Example 10.2.4

The zero vector is the vector whose initial point is also its terminal point. It is denoted by 0⃗. Its component form, in R , is 2
;
⟨0, 0⟩

in R , it is ⟨0, 0, 0⟩. Usually the context makes is clear whether 0⃗ is referring to a vector in the plane or in space.
3

Our examples have illustrated key principles in vector algebra: how to add and subtract vectors and how to multiply vectors by a
scalar. The following theorem states formally the properties of these operations.

10.2.5 https://math.libretexts.org/@go/page/4214
THEOREM 84 PROPERTIES OF VECTOR OPERATIONS

The following are true for all scalars c and d , and for all vectors u⃗, v ⃗ and w⃗ , where u⃗, v ⃗ and w⃗ are all in R or where u⃗, v ⃗ and 2

w⃗ are all in R :
3

1. u⃗ + v ⃗ = v ⃗ + u⃗

C ommutativeP roperty

2. u⃗ + v)⃗ + w⃗ = u⃗ + (v ⃗ + w⃗)

AssociativeP roperty

3. ⃗
v⃗ + 0 = v⃗

AdditiveIdentity

4. (cd)v ⃗ = c(dv)⃗
5. c(u⃗ + v)⃗ = cu⃗ + cv ⃗

DistributiveP roperty

6. (c + d)v ⃗ = cv ⃗ + dv ⃗

DistributiveP roperty

7. 0v⃗ = 0⃗
8. ∥cv⃗∥ = |c| ⋅ ∥v⃗∥
9. ∥u∥ = 0 if, and only if, u⃗ = 0⃗ .

As stated before, each vector v ⃗ conveys magnitude and direction information. We have a method of extracting the magnitude,
which we write as ∥v∥⃗ . Unit vectors are a way of extracting just the direction information from a vector.

Definition 54 Unit Vector

A unit vector is a vector v ⃗ with a magnitude of 1; that is,


⃗ = 1.
∥ v∥ (10.2.13)

Consider this scenario: you are given a vector v ⃗ and are told to create a vector of length 10 in the direction of v ⃗. How does one do
that? If we knew that u⃗ was the unit vector in the direction of v ⃗, the answer would be easy: 10u.⃗ So how do we find u⃗?
Property 8 of Theorem 84 holds the key. If we divide v ⃗ by its magnitude, it becomes a vector of length 1. Consider:
1 1 1
∥ ∥
v⃗ = ⃗
∥ v∥ (we can pull out as it is a scalar)
∥ ∥

∥ v∥ ⃗
∥ v∥ ⃗
∥ v∥

= 1.

So the vector of length 10 in the direction of v ⃗ is 10 ⃗


∥ v∥
1
v.⃗ An example will make this more clear.

Example 10.2.5: Using Unit Vectors

Let v ⃗ = ⟨3, 1⟩ and let w⃗ = ⟨1, 2, 2⟩.


1. Find the unit vector in the direction of v⃗.
2. Find the unit vector in the direction of w⃗ .
3. Find the vector in the direction of v⃗ with magnitude 5.
Solution
−−
1. We find ∥v∥⃗ = √10 . So the unit vector u⃗ in the direction of v ⃗ is
1 3 1
u⃗ = −

−v = ⟨ −−, −− ⟩. (10.2.14)
√10 √10 √10

2. We find ∥w⃗∥ = 3 , so the unit vector z ⃗ in the direction of w⃗ is


1 1 2 2
u⃗ = w⃗ = ⟨ , , ⟩. (10.2.15)
3 3 3 3

10.2.6 https://math.libretexts.org/@go/page/4214
3. To create a vector with magnitude 5 in the direction of v ⃗, we multiply the unit vector u⃗ by 5. Thus
−− −−
5 u⃗ = ⟨15/ √10, 5/ √10⟩ is the vector we seek. This is sketched in Figure 10.25.

Figure 10.25: Graphing vectors in Example 10.2.5. All vectors shown have their initial points at the origin.

The basic formation of the unit vector u⃗ in the direction of a vector v ⃗ leads to a interesting equation. It is:
1
v ⃗ = ∥ v∥
⃗ v.⃗ (10.2.16)

∥ v∥

We rewrite the equation with parentheses to make a point:

This equation illustrates the fact that a vector has both magnitude and direction, where we view a unit vector as supplying only
direction information. Identifying unit vectors with direction allows us to define parallel vectors.

Definition 55 Parallel Vectors


1. Unit vectors u⃗ and u⃗ are parallel if u⃗ = ±u⃗ .
1 2 1 2

2. Nonzero vectors v ⃗ and v ⃗ are parallel if their respective unit vectors are parallel.
1 2

It is equivalent to say that vectors v ⃗ and v ⃗ are parallel if there is a scalar c ≠ 0 such that v ⃗
1 2 1

= c v2 (see marginal note).
Note: 0⃗ is directionless; because ∥0∥ = 0 , there is no unit vector in the "direction'' of 0⃗.

Some texts define two vectors as being parallel if one is a scalar multiple of the other. By this definition, 0⃗ is parallel to all vectors
as 0⃗ = 0v ⃗ for all v ⃗.
We prefer the given definition of parallel as it is grounded in the fact that unit vectors provide direction information. One may
adopt the convention that 0⃗ is parallel to all vectors if they desire.
If one graphed all unit vectors in R with the initial point at the origin, then the terminal points would all lie on the unit circle.
2

Based on what we know from trigonometry, we can then say that the component form of all unit vectors in R is ⟨cos θ, sin θ⟩ for2

some angle θ .
A similar construction in R shows that the terminal points all lie on the unit sphere. These vectors also have a particular
3

component form, but its derivation is not as straightforward as the one for unit vectors in R . Important concepts about unit vectors
2

are given in the following Key Idea.

KEY IDEA 48 UNIT VECTORS


1. The unit vector in the direction of v ⃗ is
1
u⃗ = v.⃗ (10.2.17)

∥ v∥

2. A vector u⃗ in R is a unit vector if, and only if, its component form is ⟨cos θ, sin θ⟩ for some angle θ .
2

10.2.7 https://math.libretexts.org/@go/page/4214
3. A vector u⃗ in R is a unit vector if, and only if, its component form is ⟨sin θ cos φ, sin θ sin φ, cos θ⟩ for some angles θ and
3

φ.

These formulas can come in handy in a variety of situations, especially the formula for unit vectors in the plane.

Example 10.2.6: Finding Component Forces

Consider a weight of 50lb hanging from two chains, as shown in Figure 10.26. One chain makes an angle of 30

with the
vertical, and the other an angle of 45 . Find the force applied to each chain.

Figure 10.26: A diagram of a weight hanging from 2 chains in Example 10.2.6.


Solution
Knowing that gravity is pulling the 50lb weight straight down, we can create a vector F ⃗ to represent this force.

F = 50⟨0, −1⟩ = ⟨0, −50⟩. (10.2.18)

We can view each chain as "pulling'' the weight up, preventing it from falling. We can represent the force from each chain with
a vector. Let F ⃗ represent the force from the chain making an angle of 30 with the vertical, and let F ⃗ represent the force
1

2

form the other chain. Convert all angles to be measured from the horizontal (as shown in Figure 10.27), and apply Key Idea 48.
As we do not yet know the magnitudes of these vectors, (that is the problem at hand), we use m and m to represent them.
1 2

⃗ ∘ ∘
F 1 = m1 ⟨cos 120 , sin 120 ⟩ (10.2.19)

⃗ ∘ ∘
F 2 = m2 ⟨cos 45 , sin 45 ⟩ (10.2.20)

As the weight is not moving, we know the sum of the forces is 0⃗. This gives:
⃗ ⃗ ⃗ ⃗
F +F1 +F2 = 0

∘ ∘ ∘ ∘ ⃗
⟨0, −50⟩ + m1 ⟨cos 120 , sin 120 ⟩ + m2 ⟨cos 45 , sin 45 ⟩ = 0

Figure 10.27: A diagram of the force vectors from Example 10.2.6.


The sum of the entries in the first component is 0, and the sum of the entries in the second component is also 0. This leads us to
the following two equations:
∘ ∘
m1 cos 120 + m2 cos 45 =0
∘ ∘
m1 sin 120 + m2 sin 45 = 50

This is a simple 2-equation, 2-unkown system of linear equations. We leave it to the reader to verify that the solution is

– 50 √2
m1 = 50(√3 − 1) ≈ 36.6; m2 = – ≈ 25.88. (10.2.21)
1 + √3

10.2.8 https://math.libretexts.org/@go/page/4214
It might seem odd that the sum of the forces applied to the chains is more than 50lb. We leave it to a physics class to discuss
the full details, but offer this short explanation. Our equations were established so that the vertical components of each force
sums to 50lb, thus supporting the weight. Since the chains are at an angle, they also pull against each other, creating an
"additional'' horizontal force while holding the weight in place.

Unit vectors were very important in the previous calculation; they allowed us to define a vector in the proper direction but with an
unknown magnitude. Our computations were then computed component--wise. Because such calculations are often necessary, the
standard unit vectors can be useful.

Definition 56 STANDARD UNIT VECTORS


1. In R , the standard unit vectors are
2

i ⃗ = ⟨1, 0⟩ and j ⃗ = ⟨0, 1⟩. (10.2.22)

2. In R , the standard unit vectors are


3

⃗ ⃗ ⃗
i = ⟨1, 0, 0⟩ and j = ⟨0, 1, 0⟩ and k = ⟨0, 0, 1⟩. (10.2.23)

Example 10.2.7: Using standard unit vectors


1. Rewrite v ⃗ = ⟨2, −3⟩ using the standard unit vectors.
2. Rewrite w⃗ = 4i ⃗ − 5j ⃗ + 2k⃗ in component form.
Solution

1. v ⃗ = ⟨2, −3⟩ (10.2.24)

= ⟨2, 0⟩ + ⟨0, −3⟩ (10.2.25)

= 2⟨1, 0⟩ − 3⟨0, 1⟩ (10.2.26)

= 2i ⃗ − 3j ⃗ (10.2.27)

2. ⃗ ⃗
w⃗ = 4 i − 5 j + 2 k

(10.2.28)

= ⟨4, 0, 0⟩ + ⟨0, −5, 0⟩ + ⟨0, 0, 2⟩ (10.2.29)

= ⟨4, −5, 2⟩ (10.2.30)

These two examples demonstrate that converting between component form and the standard unit vectors is rather
straightforward. Many mathematicians prefer component form, and it is the preferred notation in this text. Many engineers
prefer using the standard unit vectors, and many engineering text use that notation.

Example 10.2.8: Finding Component Force

A weight of 25lb is suspended from a chain of length 2ft while a wind pushes the weight to the right with constant force of 5lb
as shown in Figure 10.28. What angle will the chain make with the vertical as a result of the wind's pushing? How much higher
will the weight be?

Figure 10.28: A figure of a weight being pushed by the wind in Example 10.2.8.
Solution
The force of the wind is represented by the vector F ⃗ = 5i ⃗ . The force of gravity on the weight is represented by F ⃗ = −25j ⃗ .
w g

The direction and magnitude of the vector representing the force on the chain are both unknown. We represent this force with
⃗ ⃗ ⃗
F c = m⟨cos φ, sin φ⟩ = m cos φ i + m sin φ j (10.2.31)

10.2.9 https://math.libretexts.org/@go/page/4214
for some magnitude m and some angle with the horizontal φ . (Note: θ is the angle the chain makes with the vertical; φ is the
angle with the horizontal.)
As the weight is at equilibrium, the sum of the forces is 0⃗:
⃗ ⃗ ⃗ ⃗
Fc +Fw +Fg = 0

⃗ ⃗ ⃗ ⃗ ⃗
m cos φ i + m sin φ j + 5 i − 25 j = 0

Thus the sum of the i ⃗ and j ⃗ components are 0, leading us to the following system of equations:

5 + m cos φ = 0 (10.2.32)

−25 + m sin φ = 0 (10.2.33)

This is enough to determine F ⃗ already, as we know m cos φ = −5 and m sin φ = 25 . Thus F


c c = ⟨−5, 25⟩. We can use this
to find the magnitude m:
−−−−−−−−−−
2 2 −−
m = √ (−5 ) + 25 = 5 √26 ≈ 25.5lb. (10.2.34)

We can then use either equality from Equation 10.2.33 to solve for φ . We choose the first equality as using arccosine will
return an angle in the 2 quadrant:
nd

−− −1
−5 ∘
5 + 5 √26 cos φ = 0 ⇒ φ = cos ( −− ) ≈ 1.7682 ≈ 101.31 . (10.2.35)
5 √26

Subtracting 90 from this angle gives us an angle of 11.31 with the vertical.
∘ ∘

We can now use trigonometry to find out how high the weight is lifted. The diagram shows that a right triangle is formed with
the 2ft chain as the hypotenuse with an interior angle of 11.31 . The length of the adjacent side (in the diagram, the dashed

vertical line) is 2 cos 11.31 ≈ 1.96ft. Thus the weight is lifted by about 0.04ft, almost 1/2in.

The algebra we have applied to vectors is already demonstrating itself to be very useful. There are two more fundamental
operations we can perform with vectors, the dot product and the cross product. The next two sections explore each in turn.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 10.2: An Introduction to Vectors is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory
Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

10.2.10 https://math.libretexts.org/@go/page/4214
10.3: The Dot Product
The previous section introduced vectors and described how to add them together and how to multiply them by scalars. This section
introduces a multiplication on vectors called the dot product.

Definition 57 Dot Product


1. Let u⃗ = ⟨u 1, u2 ⟩ and v ⃗ = ⟨v 1, v2 ⟩ in R . The dot product of u⃗ and v ⃗, denoted u⃗ ⋅ v ⃗ , is
2

u⃗ ⋅ v ⃗ = u1 v1 + u2 v2 . (10.3.1)

2. Let u⃗ = ⟨u 1, u2 , u3 ⟩ and v ⃗ = ⟨v 1, v2 , v3 ⟩ in R . The dot product of u⃗ and v ⃗, denoted u⃗ ⋅ v ⃗ , is


3

u⃗ ⋅ v ⃗ = u1 v1 + u2 v2 + u3 v3 . (10.3.2)

Note how this product of vectors returns a scalar, not another vector. We practice evaluating a dot product in the following
example, then we will discuss why this product is useful.

Example 10.3.1: Evaluating dot products


1. Let u⃗ = ⟨1, 2⟩ , v⃗ = ⟨3, −1⟩ in R . Find u⃗ ⋅ v⃗ .
2

2. Let x⃗ = ⟨2, −2, 5⟩ and y ⃗ = ⟨−1, 0, 3⟩ in R . Find x⃗ ⋅ y ⃗ . 3

Solution
1. Using Definition 57, we have
u⃗ ⋅ v ⃗ = 1(3) + 2(−1) = 1. (10.3.3)

2. Using the definition, we have


x⃗ ⋅ y ⃗ = 2(−1) − 2(0) + 5(3) = 13. (10.3.4)

The dot product, as shown by the preceding example, is very simple to evaluate. It is only the sum of products. While the definition
gives no hint as to why we would care about this operation, there is an amazing connection between the dot product and angles
formed by the vectors. Before stating this connection, we give a theorem stating some of the properties of the dot product.

THEOREM 85 PROPERTIES OF THE DOT PRODUCT


Let u⃗, v ⃗ and w⃗ be vectors in R or R and let c be a scalar.
2 3

1. u⃗ ⋅ v ⃗ = v ⃗ ⋅ u⃗ Commutative Property
2. u⃗ ⋅ (v ⃗ + w⃗) = u⃗ ⋅ v ⃗ + u⃗ ⋅ w⃗ Distributive Property
3. c(u⃗ ⋅ v)⃗ = (cu⃗) ⋅ v ⃗ = u⃗ ⋅ (cv)⃗
4. 0⃗ ⋅ v ⃗ = 0
5. v ⃗ ⋅ v ⃗ = ∥v∥⃗
2

The last statement of the theorem makes a handy connection between the magnitude of a vector and the dot product with itself.
Our definition and theorem give properties of the dot product, but we are still likely wondering "What does the dot product mean?''
It is helpful to understand that the dot product of a vector with itself is connected to its magnitude.
The next theorem extends this understanding by connecting the dot product to magnitudes and angles. Given vectors u⃗ and v ⃗ in the
plane, an angle θ is clearly formed when u⃗ and v ⃗ are drawn with the same initial point as illustrated in Figure 10.29(a). (We always
take θ to be the angle in [0, π] as two angles are actually created.)

10.3.1 https://math.libretexts.org/@go/page/4215
Figure 10.29: Illustrating the angle formed by two vectors with the same initial point.
The same is also true of 2 vectors in space: given u⃗ and v⃗ in R with the same initial point, there is a plane that contains both u⃗ and
3

v⃗. (When u⃗ and v⃗ are co-linear, there are infinite planes that contain both vectors.) In that plane, we can again find an angle θ

between them (and again, 0 ≤ θ ≤ π ). This is illustrated in Figure 10.29(b).


The following theorem connects this angle θ to the dot product of u⃗ and v ⃗.

theorem 86 the dot product and angles

Let u⃗ and v ⃗ be vectors in R or R . Then


2 3

u⃗ ⋅ v ⃗ = ∥ u⃗∥ ∥ v∥
⃗ cos θ, (10.3.5)

where θ , 0 ≤ θ ≤ π , is the angle between u⃗ and v ⃗.

When θ is an acute angle (i.e., 0 ≤ θ < π/2 ), cos θ is positive; when θ = π/2 , cos θ = 0 ; when θ is an obtuse angle (
π/2 < θ ≤ π ), cos θ is negative. Thus the sign of the dot product gives a general indication of the angle between the vectors,

illustrated in Figure 10.30.

Figure 10.30: Illustrating the relationship between the angle between vectors and the sign of their dot product.
We can use Theorem 86 to compute the dot product, but generally this theorem is used to find the angle between known vectors
(since the dot product is generally easy to compute). To this end, we rewrite the theorem's equation as
u⃗ ⋅ v ⃗ u⃗ ⋅ v ⃗
−1
cos θ = ⇔ θ = cos ( ). (10.3.6)
∥ u⃗∥∥ v∥
⃗ ∥ u⃗∥∥ v∥

We practice using this theorem in the following example.

Example 10.3.2: Using the dot product to find angles

Let u⃗ = ⟨3, 1⟩ , v ⃗ = ⟨−2, 6⟩ and w⃗ = ⟨−4, 3⟩ , as shown in Figure 10.31. Find the angles α , β and θ .

10.3.2 https://math.libretexts.org/@go/page/4215
Figure 10.31: Vectors used in Example 10.3.2.
Solution
We start by computing the magnitude of each vector.
−− −−
∥ u⃗∥ = √10; ⃗ = 2 √10;
∥ v∥ ∥ w⃗ ∥ = 5. (10.3.7)

We now apply Theorem 86 to find the angles.


u⃗ ⋅ v ⃗
−1
α = cos ( −− −− )
(√10)(2 √10)
π
−1 ∘
= cos (0) = = 90 .
2

−1
v ⃗ ⋅ w⃗
β = cos ( − − )
(2 √10)(5)

26
−1
= cos ( −−)
10 √10

≈ 0.6055 ≈ 34.7 .

−1
u⃗ ⋅ w⃗
θ = cos ( −− )
(√10)(5)

−9
−1
= cos ( )
−−
5 √10

≈ 2.1763 ≈ 124.7

We see from our computation that α + β = θ , as indicated by Figure 10.31. While we knew this should be the case, it is nice to see
that this non-intuitive formula indeed returns the results we expected.
We do a similar example next in the context of vectors in space.

Example 10.3.3: Using the dot product to find angles

Let u⃗ = ⟨1, 1, 1⟩, v ⃗ = ⟨−1, 3, −2⟩ and w⃗ = ⟨−5, 1, 4⟩ , as illustrated in Figure 10.32. Find the angle between each pair of
vectors.

Figure 10.32: Vectors used in Example 10.3.3

10.3.3 https://math.libretexts.org/@go/page/4215
Solution
1. Between u⃗ and v⃗:
u⃗ ⋅ v⃗
−1
θ = cos ( )
∥ u⃗∥∥ v⃗∥

0
−1
= cos ( −)
– −
√3√14
π
= .
2

2. Between u⃗ and w⃗ :
u⃗ ⋅ w⃗
−1
θ = cos ( )
∥ u⃗∥∥ w⃗ ∥

−1
0
= cos ( – −−)
√3√42
π
= .
2

3. Between v ⃗ and w⃗ :

−1
v ⃗ ⋅ w⃗
θ = cos ( )

∥ v∥∥ w⃗ ∥

0
−1
= cos ( )
−− −−
√14√42
π
= .
2

While our work shows that each angle is π/2, i.e., 90 , none of these angles looks to be a right angle in Figure 10.32. Such is

the case when drawing three--dimensional objects on the page.

All three angles between these vectors was π/2, or 90 . We know from geometry and everyday life that 90 angles are "nice'' for a
∘ ∘

variety of reasons, so it should seem significant that these angles are all π/2. Notice the common feature in each calculation (and
also the calculation of α in Example 10.3.2): the dot products of each pair of angles was 0. We use this as a basis for a definition of
the term orthogonal, which is essentially synonymous to perpendicular.

Definition 58 ORTHOGONAL
Vectors u⃗ and v ⃗ are orthogonal if their dot product is 0.

Note: The term perpendicular originally referred to lines. As mathematics progressed, the concept of "being at right angles to'' was
applied to other objects, such as vectors and planes, and the term orthogonal was introduced. It is especially used when discussing
objects that are hard, or impossible, to visualize: two vectors in 5-dimensional space are orthogonal if their dot product is 0. It is not
wrong to say they are perpendicular, but common convention gives preference to the word orthogonal.

Example 10.3.4: Finding orthogonal vectors

Let u⃗ = ⟨3, 5⟩ and v⃗ = ⟨1, 2, 3⟩.


1. Find two vectors in R that are orthogonal to u⃗.
2

2. Find two non--parallel vectors in R that are orthogonal to v ⃗.


3

Solution
1. Recall that a line perpendicular to a line with slope m has slope −1/m, the "opposite reciprocal slope.'' We can think of the
slope of u⃗ as 5/3, its "rise over run.'' A vector orthogonal to u⃗ will have slope −3/5. There are many such choices, though
all parallel:

10.3.4 https://math.libretexts.org/@go/page/4215
⟨−5, 3⟩ or ⟨5, −3⟩ or ⟨−10, 6⟩ or ⟨15, −9⟩, etc. (10.3.8)

2. There are infinite directions in space orthogonal to any given direction, so there are an infinite number of non--parallel
vectors orthogonal to v⃗. Since there are so many, we have great leeway in finding some.

One way is to arbitrarily pick values for the first two components, leaving the third unknown. For instance, let
v⃗ = ⟨2, 7, z⟩ . If v⃗ is to be orthogonal to v⃗, then v⃗ ⋅ v⃗ = 0 , so
1 1 1

−16
2 + 14 + 3z = 0 ⇒ z = . (10.3.9)
3

So v ⃗ = ⟨2, 7, −16/3⟩ is orthogonal to v ⃗. We can apply a similar technique by leaving the first or second component
1

unknown.

Another method of finding a vector orthogonal to v ⃗ mirrors what we did in part 1. Let v ⃗ = ⟨−2, 1, 0⟩ . Here we switched
2

the first two components of v ⃗, changing the sign of one of them (similar to the "opposite reciprocal'' concept before).
Letting the third component be 0 effectively ignores the third component of v ⃗, and it is easy to see that
⃗ ⋅ v ⃗ = ⟨−2, 1, 0⟩ ⋅ ⟨1, 2, 3⟩ = 0.
v2 (10.3.10)

Clearly v ⃗ and v ⃗ are not parallel.


1 2

An important construction is illustrated in Figure 10.33, where vectors u⃗ and v ⃗ are sketched. In part (a), a dotted line is drawn from
the tip of u⃗ to the line containing v ⃗, where the dotted line is orthogonal to v ⃗. In part (b), the dotted line is replaced with the vector z ⃗
and w⃗ is formed, parallel to v ⃗. It is clear by the diagram that u⃗ = w⃗ + z ⃗ . What is important about this construction is this: u⃗ is
decomposed as the sum of two vectors, one of which is parallel to v ⃗ and one that is perpendicular to v ⃗. It is hard to overstate the
importance of this construction (as we'll see in upcoming examples).
The vectors w⃗ , z ⃗ and u⃗ as shown in Figure 10.33 (b) form a right triangle, where the angle between v ⃗ and u⃗ is labeled θ . We can
find w⃗ in terms of v ⃗ and u⃗.

Figure 10.33: Developing the construction of the orthogonal projection.


Using trigonometry, we can state that
∥ w⃗ ∥ = ∥ u⃗∥ cos θ. (10.3.11)

10.3.5 https://math.libretexts.org/@go/page/4215
We also know that w⃗ is parallel to to v⃗; that is, the direction of w⃗ is the direction of v⃗, described by the unit vector 1


v⃗ . The vector
∥ v∥

w⃗ is the vector in the direction 1


∥ v∥
v⃗ with magnitude ∥u⃗∥ cos θ:

1
w⃗ = (∥ u⃗∥ cos θ) v.⃗

∥ v∥

Replace cos θ using Theorem 86:

u⃗ ⋅ v ⃗ 1
= (∥ u⃗∥ ) v⃗
∥ u⃗∥∥ v∥
⃗ ⃗
∥ v∥

u⃗ ⋅ v ⃗
= v.⃗
2

∥v∥

Now apply Theorem 85.

u⃗ ⋅ v ⃗
= v.⃗
v⃗ ⋅ v⃗

Since this construction is so important, it is given a special name.

Definition 59 orthogonal projection


Let u⃗ and v⃗ be given. The orthogonal projection of u⃗ onto v⃗, denoted proj v⃗
u⃗, is
u⃗ ⋅ v⃗
proj ⃗ u⃗ = v⃗. (10.3.12)
v
v⃗ ⋅ v⃗

Example 10.3.5: Computing the orthogonal projection

1. Let u⃗ = ⟨−2, 1⟩ and v ⃗ = ⟨3, 1⟩ . Find proj u,⃗ and sketch all three vectors with initial points at the origin.
v⃗

2. Let w⃗ = ⟨2, 1, 3⟩ and x⃗ = ⟨1, 1, 1⟩. Find proj w⃗, and sketch all three vectors with initial points at the origin.
x⃗

Figure 10.34: Graphing the vectors used in Example 10.3.5

10.3.6 https://math.libretexts.org/@go/page/4215
Solution
1. Applying Definition 59, we have
u⃗ ⋅ v⃗
proj ⃗ u⃗ = v⃗
v
v⃗ ⋅ v⃗
−5
= ⟨3, 1⟩
10
3 1
= ⟨− ,− ⟩.
2 2

Vectors u⃗, v ⃗ and proj u⃗ are sketched in Figure 10.34(a). Note how the projection is parallel to v ⃗; that is, it lies on the same
v⃗

line through the origin as v ⃗, although it points in the opposite direction. That is because the angle between u⃗ and v ⃗ is obtuse
(i.e., greater than 90 ).

2. Apply the definition:


w⃗ ⋅ x⃗
proj x⃗ w⃗ = x⃗
x⃗ ⋅ x⃗
6
= ⟨1, 1, 1⟩
3

= ⟨2, 2, 2⟩.

These vectors are sketched in Figure 10.34(b), and again in part (c) from a different perspective. Because of the nature of
graphing these vectors, the sketch in part (b) makes it difficult to recognize that the drawn projection has the geometric
properties it should. The graph shown in part (c) illustrates these properties better.

Consider Figure 10.35 where the concept of the orthogonal projection is again illustrated. It is clear that

u⃗ = proj v ⃗ u⃗ + z .⃗ (10.3.13)

Figure 10.35: Illustrating the orthogonal projection.


As we know what u⃗ and proj v⃗
u⃗ are, we can solve for z ⃗ and state that
z ⃗ = u⃗ − proj ⃗ u⃗. (10.3.14)
v

This leads us to rewrite Equation 10.3.13 in a seemingly silly way:


u⃗ = proj ⃗ u⃗ + (u⃗ − proj ⃗ u⃗). (10.3.15)
v v

This is not nonsense, as pointed out in the following Key Idea. (Notation note: the expression "∥y ''⃗ means "is parallel to y .''
⃗ We can
use this notation to state "x ∥ y '' which means "x is parallel to y .'' The expression "⊥ y '' means "is orthogonal to y ,'' and is used
⃗ ⃗ ⃗ ⃗ ⃗ ⃗

similarly.)

10.3.7 https://math.libretexts.org/@go/page/4215
key idea 49 orthogonal decomposition of vectors
Let u⃗ and v⃗ be given. Then u⃗ can be written as the sum of two vectors, one of which is parallel to v⃗, and one of which is
orthogonal to v⃗:

u⃗ = proj v ⃗ u⃗ + ( u⃗ − proj v ⃗ u⃗ ). (10.3.16)


 
∥ v⃗ ⊥ v⃗

We illustrate the use of this equality in the following example.

Example 10.3.6: Orthogonal decomposition of vectors


1. Let u⃗ = ⟨−2, 1⟩ and v ⃗ = ⟨3, 1⟩ as in Example 10.3.5. Decompose u⃗ as the sum of a vector parallel to v ⃗ and a vector
orthogonal to v ⃗.
2. Let w⃗ = ⟨2, 1, 3⟩ and x⃗ = ⟨1, 1, 1⟩ as in Example 10.3.5. Decompose w⃗ as the sum of a vector parallel to x⃗ and a vector
orthogonal to x⃗.
Solution
1. In Example 10.3.5, we found that proj v⃗
u⃗ = ⟨−1.5, −0.5⟩ . Let
z ⃗ = u⃗ − proj v ⃗ u⃗ = ⟨−2, 1⟩ − ⟨−1.5, −0.5⟩ = ⟨−0.5, 1.5⟩. (10.3.17)

Is z ⃗ orthogonal to v⃗? (I.e, is z ⃗ ⊥ v⃗ ?) We check for orthogonality with the dot product:

z ⃗ ⋅ v ⃗ = ⟨−0.5, 1.5⟩ ⋅ ⟨3, 1⟩ = 0. (10.3.18)

Since the dot product is 0, we know z ⃗ ⊥ v⃗ . Thus:

u⃗ = proj v ⃗ u⃗ + (u⃗ − proj v ⃗ u⃗)

⟨−2, 1⟩ = ⟨−1.5, −0.5⟩ + ⟨−0.5, 1.5⟩.


 
∥ v⃗ ⊥ v⃗

2. We found in Example 10.3.5 that proj x⃗


w⃗ = ⟨2, 2, 2⟩ . Applying the Key Idea, we have:
z ⃗ = w⃗ − proj ⃗ w⃗ = ⟨2, 1, 3⟩ − ⟨2, 2, 2⟩ = ⟨0, −1, 1⟩. (10.3.19)
x

We check to see if z ⃗ ⊥ x⃗ :
z ⃗ ⋅ x⃗ = ⟨0, −1, 1⟩ ⋅ ⟨1, 1, 1⟩ = 0. (10.3.20)

Since the dot product is 0, we know the two vectors are orthogonal.

We now write w⃗ as the sum of two vectors, one parallel and one orthogonal to x⃗:
w⃗ = proj ⃗ w⃗ + (w⃗ − proj ⃗ w⃗ )
x x

⟨2, 1, 3⟩ = ⟨2, 2, 2⟩ + ⟨0, −1, 1⟩


 
∥ x⃗ ⊥ x⃗

We give an example of where this decomposition is useful.

10.3.8 https://math.libretexts.org/@go/page/4215
Example 10.3.7: Orthogonally decomposing a force vector

Consider Figure 10.36(a), showing a box weighing 50lb on a ramp that rises 5ft over a span of 20ft. Find the components of
force, and their magnitudes, acting on the box (as sketched in part (b) of the figure):
1. in the direction of the ramp, and
2. orthogonal to the ramp.

Figure 10.36: Sketching the ramp and box in Example 10.3.7. Note: The vectors are not drawn to scale.
Solution
As the ramp rises 5ft over a horizontal distance of 20ft, we can represent the direction of the ramp with the vector r ⃗ = ⟨20, 5⟩.
Gravity pulls down with a force of 50lb, which we represent with g ⃗ = ⟨0, −50⟩ .
1. To find the force of gravity in the direction of the ramp, we compute proj r⃗
:
g⃗

g⃗ ⋅ r⃗
proj r ⃗ g ⃗ = r⃗
r⃗ ⋅ r⃗
−250
= ⟨20, 5⟩
425
200 50
= ⟨− ,− ⟩ ≈ ⟨−11.76, −2.94⟩.
17 17

−−
The magnitude of proj g ⃗ is ∥proj g ∥⃗ = 50/√17 ≈ 12.13lb. Though the box weighs 50lb, a force of about 12lb is enough
r⃗ r⃗

to keep the box from sliding down the ramp.


2. To find the component z ⃗ of gravity orthogonal to the ramp, we use Key Idea 49.

z ⃗ = g ⃗ − proj r ⃗ g ⃗

200 800
=⟨ ,− ⟩ ≈ ⟨11.76, −47.06⟩.
17 17

The magnitude of this force is ∥z ∥⃗ ≈ 48.51lb. In physics and engineering, knowing this force is important when computing
things like static frictional force. (For instance, we could easily compute if the static frictional force alone was enough to
keep the box from sliding down the ramp.)

Application to Work
In physics, the application of a force F to move an object in a straight line a distance d produces work; the amount of work W is
W = F d , (where F is in the direction of travel). The orthogonal projection allows us to compute work when the force is not in the

10.3.9 https://math.libretexts.org/@go/page/4215
direction of travel.

Figure 10.37: Finding work when the force and direction of travel are given as vectors.

Consider Figure 10.37, where a force F ⃗ is being applied to an object moving in the direction of d .⃗ (The distance the object travels
is the magnitude of d .)

The work done is the amount of force in the direction of d ,⃗ ∥proj F ∥⃗ , times ∥d∥: d

⃗ ⃗
F ⋅d
⃗ ∥ ∥

∥ proj ⃗F ∥ ⋅ ∥d∥ = d ⋅ ∥d∥
d ∥ ∥
⃗ ⃗
d ⋅d

∣⃗ ⃗ ∣
F ⋅d
=∣ ∣ ⋅ ∥d∥ ⋅ ∥d∥
2
∣ ∥d∥ ∣

∣ ⃗ ∣⃗
F ⋅d
∣ ∣
2
= ∥d∥
2
∥d∥

∣ ⃗ ∣⃗
= F ⋅d .
∣ ∣

The expression F ⃗ ⋅ d ⃗ will be positive if the angle between F



and d ⃗ is acute; when the angle is obtuse (hence ⃗
F ⋅d

is negative),
the force is causing motion in the opposite direction of d , resulting in "negative work.'' We want to capture this sign, so we drop the

absolute value and find that W = F ⃗ ⋅ d ⃗ .

Definition 60 WORK

Let F ⃗ be a constant force that moves an object in a straight line from point P to point Q. Let d ⃗ = P Q . The work W done by
F along d is W = F ⋅ d .
⃗ ⃗ ⃗ ⃗

Example 10.3.8: Computing work

A man slides a box along a ramp that rises 3ft over a distance of 15ft by applying 50lb of force as shown in Figure 10.38.
Compute the work done.
Solution
The figure indicates that the force applied makes a 30 angle with the horizontal, so
∘ ⃗ ∘ ∘
F = 50⟨cos 30 , sin 30 ⟩ ≈ ⟨43.3, 25⟩.

The ramp is represented by d ⃗ = ⟨15, 3⟩ . The work done is simply


⃗ ⃗ ∘ ∘
F ⋅ d = 50⟨cos 30 , sin 30 ⟩ ⋅ ⟨15, 3⟩ ≈ 724.5ft--lb. (10.3.21)

Figure 10.38: Computing work when sliding a box up a ramp in Example 10.3.8.
Note how we did not actually compute the distance the object traveled, nor the magnitude of the force in the direction of travel;
this is all inherently computed by the dot product!

The dot product is a powerful way of evaluating computations that depend on angles without actually using angles. The next
section explores another "product'' on vectors, the cross product. Once again, angles play an important role, though in a much

10.3.10 https://math.libretexts.org/@go/page/4215
different way.

This page titled 10.3: The Dot Product is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et
al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

10.3.11 https://math.libretexts.org/@go/page/4215
10.4: The Cross Product
"Orthogonality'' is immensely important. A quick scan of your current environment will undoubtedly reveal numerous surfaces and
edges that are perpendicular to each other (including the edges of this page). The dot product provides a quick test for
orthogonality: vectors u⃗ and v ⃗ are perpendicular if, and only if, u⃗ ⋅ v ⃗ = 0 .
Given two non-parallel, nonzero vectors u⃗ and v ⃗ in space, it is very useful to find a vector w⃗ that is perpendicular to both u⃗ and v ⃗.
There is a operation, called the cross product, that creates such a vector. This section defines the cross product, then explores its
properties and applications.

Definition 61 Cross product


Let u⃗ = ⟨u 1, u2 , u3 ⟩ and v ⃗ = ⟨v
1, v2 , v3 ⟩ be vectors in R . The cross product of u⃗ and v ⃗, denoted u⃗ × v ⃗ , is the vector
3

u⃗ × v⃗ = ⟨u2 v3 − u3 v2 , −(u1 v3 − u3 v1 ), u1 v2 − u2 v1 ⟩. (10.4.1)

This definition can be a bit cumbersome to remember. After an example we will give a convenient method for computing the cross
product. For now, careful examination of the products and differences given in the definition should reveal a pattern that is not too
difficult to remember. (For instance, in the first component only 2 and 3 appear as subscripts; in the second component, only 1 and
3 appear as subscripts. Further study reveals the order in which they appear.)
Let's practice using this definition by computing a cross product.

Example 10.4.1: Computing a cross product

Let u⃗ = ⟨2, −1, 4⟩ and v ⃗ = ⟨3, 2, 5⟩. Find u⃗ × v ⃗ , and verify that it is orthogonal to both u⃗ and v ⃗.

Solution
Using Definition 61, we have

u⃗ × v ⃗ = ⟨(−1)5 − (4)2, − ((2)5 − (4)3) , (2)2 − (−1)3⟩ = ⟨−13, 2, 7⟩. (10.4.2)

(We encourage the reader to compute this product on their own, then verify their result.)

We test whether or not u⃗ × v ⃗ is orthogonal to u⃗ and v ⃗ using the dot product:


(u⃗ × v⃗) ⋅ u⃗ = ⟨−13, 2, 7⟩ ⋅ ⟨2, −1, 4⟩ = 0, (10.4.3)

(u⃗ × v⃗) ⋅ v⃗ = ⟨−13, 2, 7⟩ ⋅ ⟨3, 2, 5⟩ = 0. (10.4.4)

Since both dot products are zero, u⃗ × v ⃗ is indeed orthogonal to both u⃗ and v ⃗.

A convenient method of computing the cross product starts with forming a particular 3 × 3 matrix, or rectangular array. The first
row comprises the standard unit vectors i ,⃗ j ⃗, and k⃗ . The second and third rows are the vectors u⃗ and v ⃗, respectively. Using u⃗ and v ⃗
from Example 10.4.1, we begin with:

Now repeat the first two columns after the original three:

This gives three full "upper left to lower right'' diagonals, and three full "upper right to lower left'' diagonals, as shown. Compute
the products along each diagonal, then add the products on the right and subtract the products on the left:

10.4.1 https://math.libretexts.org/@go/page/4216
⃗ ⃗ ⃗
u⃗ × v⃗ = (−5 i ⃗ + 12 j ⃗ + 4 k ) − (−3 k + 8 i ⃗ + 10 j ⃗ ) = −13 i ⃗ + 2 j ⃗ + 7 k = ⟨−13, 2, 7⟩. (10.4.5)

We practice using this method.

Example 10.4.2: Computing a cross product

Let u⃗ = ⟨1, 3, 6⟩ and v ⃗ = ⟨−1, 2, 1⟩. Compute both u⃗ × v ⃗ and v ⃗ × u⃗ .


Solution
To compute u⃗ × v ⃗ , we form the matrix as prescribed above, complete with repeated first columns:
⃗ ⃗ ⃗ ⃗ ⃗
i j k i j

1 3 6 1 3 (10.4.6)

−1 2 1 −1 2

We let the reader compute the products of the diagonals; we give the result:

⃗ ⃗ ⃗ ⃗ ⃗ ⃗
u⃗ × v ⃗ = (3 i − 6 j + 2 k ) − (−3 k + 12 i + j ) = ⟨−9, −7, 5⟩. (10.4.7)

To compute v ⃗ × u⃗ , we switch the second and third rows of the above matrix, then multiply along diagonals and subtract:
⃗ ⃗ ⃗ ⃗ ⃗
i j k i j

−1 2 1 −1 2 (10.4.8)

1 3 6 1 3

Note how with the rows being switched, the products that once appeared on the right now appear on the left, and vice--versa.
Thus the result is:

⃗ ⃗
v⃗ × u⃗ = (12 i ⃗ + j ⃗ − 3 k ) − (2 k + 3 i ⃗ − 6 j ⃗ ) = ⟨9, 7, −5⟩, (10.4.9)

which is the opposite of u⃗ × v ⃗ . We leave it to the reader to verify that each of these vectors is orthogonal to u⃗ and v ⃗.

Properties of the Cross Product


It is not coincidence that v ⃗ × u⃗ = −(u⃗ × v)⃗ in the preceding example; one can show using Definition 61 that this will always be
the case. The following theorem states several useful properties of the cross product, each of which can be verified by referring to
the definition.

THEOREM 87 PROPERTIES OF THE CROSS PRODUCT

Let u⃗, v ⃗ and w⃗ be vectors in R and let c be a scalar. The following identities hold:
3

1. u⃗ × v ⃗ = −(v ⃗ × u⃗) Anticommutative Property


2. (a) (u⃗ + v)⃗ × w⃗ = u⃗ × w⃗ + v ⃗ × w⃗ Distributive Properties
(b) u⃗ × (v ⃗ + w⃗) = u⃗ × v ⃗ + u⃗ × w⃗
3. c(u⃗ × v)⃗ = (cu⃗) × v ⃗ = u⃗ × (cv)⃗
4. (a) u⃗ × v)⃗ ⋅ u⃗ = 0 Orthogonality Properties
(b) (u⃗ × v)⃗ ⋅ v ⃗ = 0
5. u⃗ × u⃗ = 0⃗
6. u⃗ × 0⃗ = 0⃗
7. u⃗ ⋅ (v⃗ × w⃗) = (u⃗ × v⃗) ⋅ w⃗ Triple Scalar Product

10.4.2 https://math.libretexts.org/@go/page/4216
We introduced the cross product as a way to find a vector orthogonal to two given vectors, but we did not give a proof that the
construction given in Definition 61 satisfies this property. Theorem 87 asserts this property holds; we leave it as a problem in the
Exercise section to verify this.
Property 5 from the theorem is also left to the reader to prove in the Exercise section, but it reveals something more interesting than
"the cross product of a vector with itself is 0⃗.'' Let u⃗ and v⃗ be parallel vectors; that is, let there be a scalar c such that v⃗ = cu⃗ .
Consider their cross product:

u⃗ × v ⃗ = u⃗ × (c u⃗)

= c(u⃗ × u⃗) (by Property 3 of Theorem 87)

= 0⃗. (by Property 5 of Theorem 87)

We have just shown that the cross product of parallel vectors is 0⃗. This hints at something deeper. Theorem 86 related the angle
between two vectors and their dot product; there is a similar relationship relating the cross product of two vectors and the angle
between them, given by the following theorem.

THEOREM 88 THE CROSS PRODUCT AND ANGLES


Let u⃗ and v⃗ be vectors in R . Then
3

∥ u⃗ × v∥
⃗ = ∥u∥ ∥v∥ sin θ, (10.4.10)

where θ , 0 ≤ θ ≤ π , is the angle between u⃗ and v⃗.

Note: Definition 58 (through Theorem 86) defines u⃗ and v ⃗ to be orthogonal if u⃗ ⋅ v ⃗ = 0 . We could use Theorem 88 to define u⃗ and
v ⃗ are parallel if u⃗ × v ⃗ = 0 . By such a definition, 0 would be both orthogonal and parallel to every vector. Apparent paradoxes such

as this are not uncommon in mathematics and can be very useful.

Note that this theorem makes a statement about the magnitude of the cross product. When the angle between u⃗ and v ⃗ is 0 or π (i.e.,
the vectors are parallel), the magnitude of the cross product is 0. The only vector with a magnitude of 0 is 0⃗ (see Property 9 of
Theorem 84), hence the cross product of parallel vectors is 0⃗.
We demonstrate the truth of this theorem in the following example.

Example 10.4.3: The cross product and angles

Let u⃗ = ⟨1, 3, 6⟩ and v ⃗ = ⟨−1, 2, 1⟩ as in Example 10.4.2. Verify Theorem 88 by finding θ , the angle between u⃗ and v ⃗, and
the magnitude of u⃗ × v ⃗ .

Solution
We use Theorem 86 to find the angle between u⃗ and v ⃗.

−1
u⃗ ⋅ v ⃗
θ = cos ( )
∥u∥ ∥v∥

−1
11
= cos ( −− –)
√46√6

≈ 0.8471 = 48.54 .

−−−
Our work in Example 10.4.2 showed that u⃗ × v ⃗ = ⟨−9, −7, 5⟩ , hence ∥u⃗ × v∥⃗ = √155. Is ∥u⃗ × v∥⃗ = ∥u∥ ∥v∥ sin θ ? Using
numerical approximations, we find:
−−− −− –
∥ u⃗ × v∥
⃗ = √155 ∥u∥ ∥v∥ sin θ = √46√6 sin 0.8471

≈ 12.45. ≈ 12.45.

Numerically, they seem equal. Using a right triangle, one can show that
−−−
11 √155
−1
sin( cos ( −− – )) = −− –, (10.4.11)
√46√6 √46√6

10.4.3 https://math.libretexts.org/@go/page/4216
|which allows us to verify the theorem exactly.

Right Hand Rule


The anticommutative property of the cross product demonstrates that u⃗ × v ⃗ and v ⃗ × u⃗ differ only by a sign -- these vectors have
the same magnitude but point in the opposite direction. When seeking a vector perpendicular to u⃗ and v ⃗, we essentially have two
directions to choose from, one in the direction of u⃗ × v ⃗ and one in the direction of v ⃗ × u⃗ . Does it matter which we choose? How
can we tell which one we will get without graphing, etc.?
Another wonderful property of the cross product, as defined, is that it follows the right hand rule. Given u⃗ and v ⃗ in R with the3

same initial point, point the index finger of your right hand in the direction of u⃗ and let your middle finger point in the direction of
v ⃗ (much as we did when establishing the right hand rule for the 3-dimensional coordinate system). Your thumb will naturally

extend in the direction of u⃗ × v ⃗ . One can "practice'' this using Figure 10.39. If you switch, and point the index finder in the
direction of v ⃗ and the middle finger in the direction of u⃗, your thumb will now point in the opposite direction, allowing you to
"visualize'' the anticommutative property of the cross product.

Figure 10.39: Illustrating the Right Hand Rule of the cross product.

Applications of the Cross Product


There are a number of ways in which the cross product is useful in mathematics, physics and other areas of science beyond "just''
finding a vector perpendicular to two others. We highlight a few here.
Area of a Parallelogram
It is a standard geometry fact that the area of a parallelogram is A = bh , where b is the length of the base and h is the height of the
parallelogram, as illustrated in Figure 10.40(a). As shown when defining the Parallelogram Law of vector addition, two vectors u⃗
and v ⃗ define a parallelogram when drawn from the same initial point, as illustrated in Figure 10.40(b). Trigonometry tells us that
h = ∥u∥ sin θ , hence the area of the parallelogram is

A = ∥u∥ ∥v∥ sin θ = ∥ u⃗ × v∥,


⃗ (10.4.12)

where the second equality comes from Theorem 88.

10.4.4 https://math.libretexts.org/@go/page/4216
Figure 10.40: Using the cross product to find the area of a parallelogram.
We illustrate using Equation (10.4.12) in the following example.

Example 10.4.4: Finding the area of a parallelogram

1. Find the area of the parallelogram defined by the vectors u⃗ = ⟨2, 1⟩ and v ⃗ = ⟨1, 3⟩ .
2. Verify that the points A = (1, 1, 1), B = (2, 3, 2), C = (4, 5, 3) and D = (3, 3, 2) are the vertices of a parallelogram. Find
the area of the parallelogram.
Solution
1. Figure 10.41(a) sketches the parallelogram defined by the vectors u⃗ and v⃗. We have a slight problem in that our vectors
exist in R , not R , and the cross product is only defined on vectors in R . We skirt this issue by viewing u⃗ and v ⃗ as
2 3 3

vectors in the x − y plane of R , and rewrite them as u⃗ = ⟨2, 1, 0⟩ and v ⃗ = ⟨1, 3, 0⟩. We can now compute the cross
3

product. It is easy to show that u⃗ × v ⃗ = ⟨0, 0, 5⟩ ; therefore the area of the parallelogram is A = ∥u⃗ × v∥⃗ = 5 .
2. To show that the quadrilateral ABC D is a parallelogram (shown in Figure 10.41(b)), we need to show that the opposite
→ → → →
sides are parallel. We can quickly show that AB = DC = ⟨1, 2, 1⟩ and BC = AD = ⟨2, 2, 1⟩ . We find the area by
→ →
computing the magnitude of the cross product of AB and BC :
→ → → →

AB × BC = ⟨0, 1, −2⟩ ⇒ ∥ AB × BC ∥ = √5 ≈ 2.236. (10.4.13)

10.4.5 https://math.libretexts.org/@go/page/4216
Figure 10.41: Sketching the parallelograms in Example 10.4.4.

This application is perhaps more useful in finding the area of a triangle (in short, triangles are used more often than
parallelograms). We illustrate this in the following example.

Example 10.4.5: Area of a triangle

Find the area of the triangle with vertices A = (1, 2) , B = (2, 3) and C = (3, 1) , as pictured in Figure 10.42.

Solution
We found the area of this triangle in a previous example to be 1.5 using integration. There we discussed the fact that finding
the area of a triangle can be inconvenient using the " bh '' formula as one has to compute the height, which generally involves
1

finding angles, etc. Using a cross product is much more direct.

Figure 10.42: Finding the area of a triangle in Example 10.4.5


→ →
We can choose any two sides of the triangle to use to form vectors; we choose AB = ⟨1, 1⟩ and AC = ⟨2, −1⟩ . As in the
previous example, we will rewrite these vectors with a third component of 0 so that we can apply the cross product. The area of
the triangle is
1 → → 1 1 3
∥ AB × AC ∥ = ∥⟨1, 1, 0⟩ × ⟨2, −1, 0⟩∥ = ∥⟨0, 0, −3⟩∥ = . (10.4.14)
2 2 2 2

We arrive at the same answer as before with less work.

Volume of a Parallelepiped

10.4.6 https://math.libretexts.org/@go/page/4216
The three dimensional analogue to the parallelogram is the parallelepiped. Each face is parallel to the face opposite face, as
illustrated in Figure 10.43. By crossing v ⃗ and w⃗ , one gets a vector whose magnitude is the area of the base. Dotting this vector with
u⃗ computes the volume of parallelepiped! (Up to a sign; take the absolute value.)

Figure 10.43: A parallelepiped is the three dimensional analogue to the parallelogram.


Thus the volume of a parallelepiped defined by vectors u⃗, v ⃗ and w⃗ is
V = | u⃗ ⋅ (v⃗ × w⃗ )|. (10.4.15)

Note how this is the Triple Scalar Product, first seen in Theorem 87. Applying the identities given in the theorem shows that we can
apply the Triple Scalar Product in any "order'' we choose to find the volume. That is,
V = | u⃗ ⋅ (v ⃗ × w⃗ )| = | u⃗ ⋅ (w⃗ × v)|
⃗ = |(u⃗ × v)
⃗ ⋅ w⃗ |, etc. (10.4.16)

Example 10.4.6: Finding the volume of parallelepiped

Find the volume of the parallepiped defined by the vectors u⃗ = ⟨1, 1, 0⟩, v⃗ = ⟨−1, 1, 0⟩ and w⃗ = ⟨0, 1, 1⟩.
Solution
We apply Equation (10.4.15). We first find v ⃗ × w⃗ = ⟨1, 1, −1⟩ . Then

| u⃗ ⋅ (v ⃗ × w⃗ )| = |⟨1, 1, 0⟩ ⋅ ⟨1, 1, −1⟩| = 2. (10.4.17)

So the volume of the parallelepiped is 2 cubic units.

Figure 10.44: A parallelepiped in Example 10.4.6

While this application of the Triple Scalar Product is interesting, it is not used all that often: parallelepipeds are not a common
shape in physics and engineering. The last application of the cross product is very applicable in engineering.

Torque
Torque is a measure of the turning force applied to an object. A classic scenario involving torque is the application of a wrench to a
bolt. When a force is applied to the wrench, the bolt turns. When we represent the force and wrench with vectors F ⃗ and ℓ ⃗, we see
that the bolt moves (because of the threads) in a direction orthogonal to F ⃗ and ℓ ⃗. Torque is usually represented by the Greek letter
τ , or tau, and has units of N⋅m, a Newton--meter, or ft⋅lb, a foot--pound.

10.4.7 https://math.libretexts.org/@go/page/4216
While a full understanding of torque is beyond the purposes of this book, when a force F ⃗ is applied to a lever arm ℓ ⃗, the resulting
torque is
⃗ ⃗
τ ⃗ = ℓ ×F. (10.4.18)

Example 10.4.7: Computing torque

A lever of length 2ft makes an angle with the horizontal of 45 . Find the resulting torque when a force of 10lb is applied to the

end of the level where:


1. the force is perpendicular to the lever, and
2. the force makes an angle of 60 with the lever, as shown in Figure 10.45.

Figure 10.45: Showing a force being applied to a lever in Example 10.4.7.


Solution
1. We start by determining vectors for the force and lever arm. Since the lever arm makes a 45 angle with the horizontal and

– –
is 2ft long, we can state that ⃗ ∘ ∘
ℓ = 2⟨cos 45 , sin 45 ⟩ = ⟨√2, √2⟩.

Since the force vector is perpendicular to the lever arm (as seen in the left hand side of Figure 10.45), we can conclude it is
making an angle of −45 with the horizontal. As it has a magnitude of 10lb, we can state

⃗ ∘ ∘ – –
F = 10⟨cos(−45 ), sin(−45 )⟩ = ⟨5 √2, −5 √2⟩.

Using Equation 10.45 to find the torque requires a cross product. We again let the third component of each vector be 0 and
compute the cross product:
⃗ ⃗
τ ⃗ = ℓ ×F
– – – –
= ⟨√2, √2, 0⟩ × ⟨5 √2, −5 √2, 0⟩

= ⟨0, 0, −20⟩

This clearly has a magnitude of 20 ft-lb.

We can view the force and lever arm vectors as lying "on the page''; our computation of τ ⃗ shows that the torque goes "into
the page.'' This follows the Right Hand Rule of the cross product, and it also matches well with the example of the wrench
turning the bolt. Turning a bolt clockwise moves it in.
– –
2. Our lever arm can still be represented by ℓ ⃗ = ⟨√2, √2⟩ . As our force vector makes a 60 angle with ℓ ⃗, we can see

(referencing the right hand side of the figure) that F ⃗ makes a −15 angle with the horizontal. Thus

– –
5(1 + √3) 5(1 + √3)
⃗ ∘ ∘
F = 10⟨cos −15 , sin −15 ⟩ = ⟨ – ,− – ⟩
√2 √2

≈ ⟨9.659, −2.588⟩.

We again make the third component 0 and take the cross product to find the torque:

10.4.8 https://math.libretexts.org/@go/page/4216
⃗ ⃗
τ ⃗ = ℓ ×F
– –
– – 5(1 + √3) 5(1 + √3)
= ⟨√2, √2, 0⟩ × ⟨ ,− , 0⟩
– –
√2 √2

= ⟨0, 0, −10 √3⟩

≈ ⟨0, 0, −17.321⟩.

As one might expect, when the force and lever arm vectors are orthogonal, the magnitude of force is greater than when the
vectors are not orthogonal.

While the cross product has a variety of applications (as noted in this chapter), its fundamental use is finding a vector perpendicular
to two others. Knowing a vector is orthogonal to two others is of incredible importance, as it allows us to find the equations of lines
and planes in a variety of contexts. The importance of the cross product, in some sense, relies on the importance of lines and
planes, which see widespread use throughout engineering, physics and mathematics. We study lines and planes in the next two
sections.

This page titled 10.4: The Cross Product is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman
et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

10.4.9 https://math.libretexts.org/@go/page/4216
10.5: Lines
To find the equation of a line in the x-y plane, we need two pieces of information: a point and the slope. The slope conveys
direction information. As vertical lines have an undefined slope, the following statement is more accurate:

To define a line, one needs a point on the line and the direction of the line.
This holds true for lines in space.

Let P be a point in space, let p ⃗ be the vector with initial point at the origin and terminal point at P (i.e., p ⃗ "points'' to P ), and let d ⃗
be a vector. Consider the points on the line through P in the direction of d .⃗
Clearly one point on the line is P ; we can say that the vector p ⃗ lies at this point on the line. To find another point on the line, we
can start at p ⃗ and move in a direction parallel to d .⃗ For instance, starting at p ⃗ and traveling one length of d ⃗ places one at another
point on the line. Consider Figure 10.47 where certain points along the line are indicated.

Figure 10.47: Defining a line in space.


The figure illustrates how every point on the line can be obtained by starting with p ⃗ and moving a certain distance in the direction
of d .⃗ That is, we can define the line as a function of t :
⃗ ⃗
ℓ (t) = p ⃗ + td . (10.5.1)

In many ways, this is not a new concept. Compare Equation 10.5.1 to the familiar "y = mx + b " equation of a line:

Figure 10.46: Understanding the vector equation of a line.


The equations exhibit the same structure: they give a starting point, define a direction, and state how far in that direction to travel.
Equation 10.5.1 is an example of a vector-valued function; the input of the function is a real number and the output is a vector.
We will cover vector--valued functions extensively in the next chapter.

There are other ways to represent a line. Let p ⃗ = ⟨x 0, y0 , z0 ⟩ and let d ⃗ = ⟨a, b, c⟩ . Then the equation of the line through p ⃗ in the
direction of d ⃗ is:
⃗ ⃗
ℓ (t) = p ⃗ + td

= ⟨x0 , y0 , z0 ⟩ + t⟨a, b, c⟩

= ⟨x0 + at, y0 + bt, z0 + ct⟩.

The last line states that the x values of the line are given by x = x + at , the y values are given by y = y + bt , and the z values
0 0

are given by z = z + ct . These three equations, taken together, are the parametric equations of the line through p ⃗ in the
0

10.5.1 https://math.libretexts.org/@go/page/4217
direction of d .⃗
Finally, each of the equations for x, y and z above contain the variable t . We can solve for t in each equation:
x − x0
x = x0 + at ⇒ t = ,
a
y − y0
y = y0 + bt ⇒ t = ,
b
z − z0
z = z0 + ct ⇒ t = ,
c

assuming a, b, c ≠ 0 .
Since t is equal to each expression on the right, we can set these equal to each other, forming the symmetric equations of the line
through p ⃗ in the direction of d :⃗
x − x0 y − y0 z − z0
= = . (10.5.2)
a b c

Each representation has its own advantages, depending on the context. We summarize these three forms in the following definition,
then give examples of their use.

Definition 62 Equations of Lines in Space

Consider the line in space that passes through p ⃗ = ⟨x 0, y0 , z0 ⟩ in the direction of d ⃗ = ⟨a, b, c⟩.
1. The vector equation of the line is
⃗ ⃗
ℓ (t) = p ⃗ + td . (10.5.3)

2. The parametric equations of the line are

x = x0 + at, y = y0 + bt, z = z0 + ct. (10.5.4)

3. The symmetric equations of the line are


x − x0 y − y0 z − z0
= = . (10.5.5)
a b c

Example 10.5.1: Finding the equation of a line

Give all three equations, as given in Definition 62, of the line through P = (2, 3, 1) in the direction of ⃗
d = ⟨−1, 1, 2⟩ . Does
the point Q = (−1, 6, 6) lie on this line?
Solution
We identify the point P = (2, 3, 1) with the vector p ⃗ = ⟨2, 3, 1⟩. Following the definition, we have

1. the vector equation of the line is ℓ (t)



= ⟨2, 3, 1⟩ + t⟨−1, 1, 2⟩ ;

2. the parametric equations of the line are


x = 2 − t, y = 3 + t, z = 1 + 2t; and (10.5.6)

3. the symmetric equations of the line are


x −2 y −3 z−1
= = . (10.5.7)
−1 1 2

10.5.2 https://math.libretexts.org/@go/page/4217
Figure 10.48: Graphing a line in Example 10.5.1.
The first two equations of the line are useful when a t value is given: one can immediately find the corresponding point on the
line. These forms are good when calculating with a computer; most software programs easily handle equations in these
formats. (For instance, to make Figure 10.48, a certain graphics program was given the input (2-x,3+x,1+2*x) . This
particular program requires the variable always be "x" instead of "t ").
Does the point Q = (−1, 6, 6) lie on the line? The graph in Figure 10.48 makes it clear that it does not. We can answer this
question without the graph using any of the three equation forms. Of the three, the symmetric equations are probably best
suited for this task. Simply plug in the values of x, y and z and see if equality is maintained:
−1 − 2 ? 6 −3 ? 6 −1
= = ⇒ 3 = 3 ≠ 2.5. (10.5.8)
−1 1 2

We see that Q does not lie on the line as it did not satisfy the symmetric equations.

Example 10.5.2: Finding the equation of a line through two points

Find the parametric equations of the line through the points P = (2, −1, 2) and Q = (1, 3, −1).
Solution
Recall the statement made at the beginning of this section: to find the equation of a line, we need a point and a direction. We
have two points; either one will suffice. The direction of the line can be found by the vector with initial point P and terminal

point Q: P Q = ⟨−1, 4, −3⟩.

The parametric equations of the line ℓ through P in the direction of P Q are:
ℓ : x = 2 −t y = −1 + 4t z = 2 − 3t. (10.5.9)

Figure 10.49: A graph of the line in Example 10.5.2.


A graph of the points and line are given in Figure 10.49. Note how in the given parametrization of the line, t = 0 corresponds
to the point P , and t = 1 corresponds to the point Q. This relates to the understanding of the vector equation of a line

described in Figure 10.46. The parametric equations "start'' at the point P , and t determines how far in the direction of P Q to
→ →
travel. When t = 0 , we travel 0 lengths of P Q; when t = 1 , we travel one length of P Q, resulting in the point Q.

10.5.3 https://math.libretexts.org/@go/page/4217
Parallel, Intersecting and Skew Lines
In the plane, two distinct lines can either be parallel or they will intersect at exactly one point. In space, given equations of two
lines, it can sometimes be difficult to tell whether the lines are distinct or not (i.e., the same line can be represented in different
ways). Given lines ℓ ⃗ (t) = p ⃗ + td ⃗ and ℓ ⃗ (t) = p ⃗ + td ⃗ , we have four possibilities: ℓ ⃗ and ℓ ⃗ are
1 1 1 2 2 2 1 2

The next two examples investigate these possibilities.

Example 10.5.3: Comparing lines

Consider lines ℓ and ℓ , given in parametric equation form:


1 2

x = 1 + 3t x = −2 + 4s

ℓ1 : y = 2 −t ℓ2 : y = 3 +s (10.5.10)

z = t z = 5 + 2s.

Determine whether ℓ and ℓ are the same line, intersect, are parallel, or skew.
1 2

Solution
We start by looking at the directions of each line. Line ℓ1 has the direction given by d

1 = ⟨3, −1, 1⟩ and line ℓ2 has the
direction given by d = ⟨4, 1, 2⟩. It should be clear that d and d are not parallel, hence ℓ and ℓ are not the same line, nor

2

1

2 1 2

are they parallel. Figure 10.50 verifies this fact (where the points and directions indicated by the equations of each line are
identified).

Figure 10.50: Sketching the lines from Example 10.5.3


We next check to see if they intersect (if they do not, they are skew lines). To find if they intersect, we look for t and s values
such that the respective x, y and z values are the same. That is, we want s and t such that:

1 + 3t = −2 + 4s

2 −t = 3 +s (10.5.11)

t = 5 + 2s.

This is a relatively simple system of linear equations. Since the last equation is already solved for t , substitute that value of t
into the equation above it:

2 − (5 + 2s) = 3 + s ⇒ s = −2, t = 1. (10.5.12)

A key to remember is that we have three equations; we need to check if s = −2, t =1 satisfies the first equation as well:

1 + 3(1) ≠ −2 + 4(−2). (10.5.13)

It does not. Therefore, we conclude that the lines ℓ and ℓ are skew.
1 2

10.5.4 https://math.libretexts.org/@go/page/4217
Example 10.5.4: Comparing lines

Consider lines ℓ and ℓ , given in parametric equation form:


1 2

x = −0.7 + 1.6t x = 2.8 − 2.9s

ℓ1 : y = 4.2 + 2.72t ℓ2 : y = 10.15 − 4.93s (10.5.14)

z = 2.3 − 3.36t z = −5.05 + 6.09s.

Determine whether ℓ and ℓ are the same line, intersect, are parallel, or skew.
1 2

Solution
It is obviously very difficult to simply look at these equations and discern anything. This is done intentionally. In the "real
world,'' most equations that are used do not have nice, integer coefficients. Rather, there are lots of digits after the decimal and
the equations can look "messy.''
We again start by deciding whether or not each line has the same direction. The direction of ℓ1 is given by
d

1 and the direction of ℓ is given by d ⃗ = ⟨−2.9, −4.93, 6.09⟩. When it is not clear through
= ⟨1.6, 2.72, −3.36⟩ 2 2

observation whether two vectors are parallel or not, the standard way of determining this is by comparing their respective unit
vectors. Using a calculator, we find:

d 1
u⃗1 = = ⟨0.3471, 0.5901, −0.7289⟩

∥d 1∥


d 2
u⃗2 = = ⟨−0.3471, −0.5901, 0.7289⟩.

∥d 2∥

The two vectors seem to be parallel (at least, their components are equal to 4 decimal places). In most situations, it would
suffice to conclude that the lines are at least parallel, if not the same. One way to be sure is to rewrite d ⃗ and d ⃗ in terms of 1 2

fractions, not decimals. We have


16 272 336 29 493 609
⃗ ⃗
d 1 =⟨ , ,− ⟩ d 2 = ⟨− ,− , ⟩. (10.5.15)
10 100 100 10 100 100

One can then find the magnitudes of each vector in terms of fractions, then compute the unit vectors likewise. After a lot of
manual arithmetic (or after briefly using a computer algebra system), one finds that
−−
− −−

10 17 21 10 17 21
u⃗1 = ⟨√ , −−−,− −−−⟩ u⃗2 = ⟨−√ ,− −−−, −−− ⟩. (10.5.16)
83 √830 √830 83 √830 √830

We can now say without equivocation that these lines are parallel.
Are they the same line? The parametric equations for a line describe one point that lies on the line, so we know that the point
P = (−0.7, 4.2, 2.3) lies on ℓ . To determine if this point also lies on ℓ , plug in the x, y and z values of P
1 1 into the
2 1

symmetric equations for ℓ : 2

(−0.7) − 2.8 ? (4.2) − 10.15 ? (2.3) − (−5.05)


= = ⇒ 1.2069 = 1.2069 = 1.2069. (10.5.17)
−2.9 −4.93 6.09

Figure 10.51: Graphing the lines in Example 10.5.4

10.5.5 https://math.libretexts.org/@go/page/4217
The point P lies on both lines, so we conclude they are the same line, just parametrized differently. Figure 10.51 graphs this
1

line along with the points and vectors described by the parametric equations. Note how d ⃗ and d ⃗ are parallel, though point in1 2

opposite directions (as indicated by their unit vectors above).

Distances
Given a point Q and a line ℓ (t)

= p ⃗ + td

in space, it is often useful to know the distance from the point to the line. (Here we use
the standard definition of "distance,'' i.e., the length of the shortest line segment from the point to the line.) Identifying p ⃗ with the
point P , Figure 10.52 will help establish a general method of computing this distance h .

Figure 10.52: Establishing the distance from a point to a line.



From trigonometry, we know h = ∥ P Q∥ sin θ . We have a similar identity involving the cross product:
→ →

∥ P Q × d ∥ = ∥ P Q∥ ∥d∥ sin θ. Divide both sides of this latter equation by ∥d∥ to obtain h :


∥P Q × d ∥
h = . (10.5.18)
∥d∥

Figure 10.53: Establishing the distance between lines.


It is also useful to determine the distance between lines, which we define as the length of the shortest line segment that connects the
two lines (an argument from geometry shows that this line segments is perpendicular to both lines). Let lines ℓ ⃗ (t) = p ⃗ + td ⃗ 1 1 1

and ℓ ⃗2 (t)
⃗ + td
= p2

2 be given, as shown in Figure 10.53. To find the direction orthogonal to both d

1 and d ⃗ , we take the cross
2


product: c ⃗ = d ⃗ 1 ×d

2 . The magnitude of the orthogonal projection of P 1 P2 onto c ⃗ is the distance h we seek:

h = ∥proj c⃗
P1 P2 ∥


P1 P2 ⋅ c ⃗
=∥ ⃗
c∥
c⃗ ⋅ c⃗

| P1 P2 ⋅ c |⃗
= ∥c∥
2
∥c∥


| P1 P2 ⋅ c |⃗
= .
∥c∥

A problem in the Exercise section is to show that this distance is 0 when the lines intersect. Note the use of the Triple Scalar
→ →
Product: P 1 P2 ⋅ c = P1 P2 ⋅ (d

1 ×d

2 ).

10.5.6 https://math.libretexts.org/@go/page/4217
The following Key Idea restates these two distance formulas.

KEY IDEA 50 DISTANCE TO LINES

1. Let P be a point on a line ℓ that is parallel to d .⃗ The distance h from a point Q to the line ℓ is:


∥P Q × d ∥
h = . (10.5.19)
∥d∥

2. Let P be a point on line ℓ that is parallel to d ⃗ , and let P be a point on line ℓ parallel to d ⃗ , and let c ⃗ = d ⃗
1 1 1 2 2 2 1 ×d

2 ,
where lines ℓ and ℓ are not parallel. The distance h between the two lines is:
1 2


| P1 P2 ⋅ c |⃗
h = . (10.5.20)
∥c∥

Example 10.5.5: Finding the distance from a point to a line

Find the distance from the point Q = (1, 1, 3) to the line ℓ (t)

= ⟨1, −1, 1⟩ + t⟨2, 3, 1⟩.

Solution

The equation of the line gives us the point P = (1, −1, 1) that lies on the line, hence P Q = ⟨0, 2, 2⟩. The equation also gives
d = ⟨2, 3, 1⟩ . Following Key Idea 50, we have the distance as



∥P Q × d ∥
h =
∥d∥

∥⟨−4, 4, −4⟩∥
= −−
√14

4 √3
= −− ≈ 1.852.
√14

The point Q is approximately 1.852 units from the line ℓ (t)



.

Example 10.5.6: Finding the distance between lines

Find the distance between the lines

x = 1 + 3t x = −2 + 4s

ℓ1 : y = 2 −t ℓ2 : y = 3 +s (10.5.21)

z = t z = 5 + 2s.

Solution
These are the sames lines as given in Example 10.5.3, where we showed them to be skew. The equations allow us to identify
the following points and vectors:

P1 = (1, 2, 0) P2 = (−2, 3, 5) ⇒ P1 P2 = ⟨−3, 1, 5⟩. (10.5.22)

⃗ ⃗ ⃗ ⃗
d 1 = ⟨3, −1, 1⟩ d 2 = ⟨4, 1, 2⟩ ⇒ c⃗ = d 1 ×d 2 = ⟨−3, −2, 7⟩. (10.5.23)

From Key Idea 50 we have the distance h between the two lines is

10.5.7 https://math.libretexts.org/@go/page/4217

| P1 P2 ⋅ c |⃗
h =
∥c∥

42
= ≈ 5.334.
−−
√62

The lines are approximately 5.334 units apart.

One of the key points to understand from this section is this: to describe a line, we need a point and a direction. Whenever a
problem is posed concerning a line, one needs to take whatever information is offered and glean point and direction information.
Many questions can be asked (and are asked in the Exercise section) whose answer immediately follows from this understanding.
Lines are one of two fundamental objects of study in space. The other fundamental object is the plane, which we study in detail in
the next section. Many complex three dimensional objects are studied by approximating their surfaces with lines and planes.

This page titled 10.5: Lines is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et al. via
source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

10.5.8 https://math.libretexts.org/@go/page/4217
10.6: Planes
Any flat surface, such as a wall, table top or stiff piece of cardboard can be thought of as representing part of a plane. Consider a
piece of cardboard with a point P marked on it. One can take a nail and stick it into the cardboard at P such that the nail is
perpendicular to the cardboard; see Figure 10.54.

Figure 10.54: Illustrating defining a plane with a sheet of cardboard and a nail.
This nail provides a "handle'' for the cardboard. Moving the cardboard around moves P to different locations in space. Tilting the
nail (but keeping P fixed) tilts the cardboard. Both moving and tilting the cardboard defines a different plane in space. In fact, we
can define a plane by: 1) the location of P in space, and 2) the direction of the nail.
The previous section showed that one can define a line given a point on the line and the direction of the line (usually given by a
vector). One can make a similar statement about planes: we can define a plane in space given a point on the plane and the direction
the plane "faces'' (using the description above, the direction of the nail). Once again, the direction information will be supplied by a
vector, called a normal vector, that is orthogonal to the plane.

What exactly does "orthogonal to the plane'' mean? Choose any two points P and Q in the plane, and consider the vector P Q. We
→ →
say a vector n⃗ is orthogonal to the plane if n⃗ is perpendicular to PQ for all choices of P and Q; that is, if n⃗ ⋅ P Q = 0 for all P

and Q.
This gives us way of writing an equation describing the plane. Let P = (x0 , y0 , z0 ) be a point in the plane and let n⃗ = ⟨a, b, c⟩ be

a normal vector to the plane. A point Q = (x, y, z) lies in the plane defined by P and n⃗ if, and only if, PQ is orthogonal to n⃗ .

Knowing P Q = ⟨x − x 0, y − y0 , z − z0 ⟩ , consider:

P Q ⋅ n⃗ = 0

⟨x − x0 , y − y0 , z − z0 ⟩ ⋅ ⟨a, b, c⟩ =0

a(x − x0 ) + b(y − y0 ) + c(z − z0 ) =0 (10.6.1)

Equation 10.6.1 defines an implicit function describing the plane. More algebra produces:

ax + by + cz = ax0 + b y0 + c z0 .

The right hand side is just a number, so we replace it with d :

ax + by + cz = d. (10.6.2)

As long as c ≠ 0 , we can solve for z :


1
z = (d − ax − by). (10.6.3)
c

Equation 10.6.3 is especially useful as many computer programs can graph functions in this form. Equations 10.6.1 and 10.6.2

have specific names, given next.

10.6.1 https://math.libretexts.org/@go/page/4218
Definition 63 Equations of a Plane in Standard and General Forms

a(x − x0 ) + b(y − y0 ) + c(z − z0 ) = 0; (10.6.4)

the equation's general form is


ax + by + cz = d. (10.6.5)

A key to remember throughout this section is this: to find the equation of a plane, we need a point and a normal vector. We will
give several examples of finding the equation of a plane, and in each one different types of information are given. In each case, we
need to use the given information to find a point on the plane and a normal vector.

Example 10.6.1: Finding the equation of a plane.

Write the equation of the plane that passes through the points P = (1, 1, 0) , Q = (1, 2, −1) and R = (0, 1, 2) in standard
form.

Solution
→ →
We need a vector n⃗ that is orthogonal to the plane. Since P , Q and R are in the plane, so are the vectors PQ and PR ;
→ → → →
PQ×PR is orthogonal to P Q and P R and hence the plane itself.
→ →
It is straightforward to compute n⃗ = P Q × P R = ⟨2, 1, 1⟩ . We can use any point we wish in the plane (any of P , Q or R will
do) and we arbitrarily choose P . Following Definition 63, the equation of the plane in standard form is
2(x − 1) + (y − 1) + z = 0. (10.6.6)

The plane is sketched in Figure 10.55.

Figure 10.55: Sketching the plane in Example 10.6.1

We have just demonstrated the fact that any three non-collinear points define a plane. (This is why a three-legged stool does not
"rock;'' it's three feet always lie in a plane. A four-legged stool will rock unless all four feet lie in the same plane.)\\

Example 10.6.2: Finding the equation of a plane.

Verify that lines ℓ and ℓ , whose parametric equations are given below, intersect, then give the equation of the plane that
1 2

contains these two lines in general form.


x = −5 + 2s x = 2 + 3t

ℓ1 : y = 1 +s ℓ2 : y = 1 − 2t (10.6.7)

z = −4 + 2s z = 1 +t

10.6.2 https://math.libretexts.org/@go/page/4218
Solution
The lines clearly are not parallel. If they do not intersect, they are skew, meaning there is not a plane that contains them both. If
they do intersect, there is such a plane.
To find their point of intersection, we set the x, y and z equations equal to each other and solve for s and t :

−5 + 2s = 2 + 3t

1 +s = 1 − 2t ⇒ s = 2, t = −1. (10.6.8)

−4 + 2s = 1 +t

When s = 2 and t = −1 , the lines intersect at the point P = (−1, 3, 0) .

Let d

1 = ⟨2, 1, 2⟩ and d

2 = ⟨3, −2, 1⟩ be the directions of lines ℓ1 and ℓ2 , respectively. A normal vector to the plane
containing these the two lines will also be orthogonal to d

1 and d

2 . Thus we find a normal vector n⃗ by computing
= ⟨5, 4 − 7⟩ .
⃗ ⃗
n⃗ = d × d 1 2

We can pick any point in the plane with which to write our equation; each line gives us infinite choices of points. We choose
P , the point of intersection. We follow Definition 63 to write the plane's equation in general form:

5(x + 1) + 4(y − 3) − 7z =0

5x + 5 + 4y − 12 − 7z =0

5x + 4y − 7z = 7.

The plane's equation in general form is 5x + 4y − 7z = 7 ; it is sketched in Figure 10.56.

Figure 10.56: Sketching the plane in Example 10.6.2

Example 10.6.3: Finding the equation of a plane

Give the equation, in standard form, of the plane that passes through the point P = (−1, 0, 1) and is orthogonal to the line
with vector equation ⃗
ℓ (t) = ⟨−1, 0, 1⟩ + t⟨1, 2, 2⟩ .
Solution

As the plane is to be orthogonal to the line, the plane must be orthogonal to the direction of the line given by d ⃗ = ⟨1, 2, 2⟩ . We
use this as our normal vector. Thus the plane's equation, in standard form, is
(x + 1) + 2y + 2(z − 1) = 0. (10.6.9)

The line and plane are sketched in Figure 10.57.

10.6.3 https://math.libretexts.org/@go/page/4218
Figure 10.57: The line and plane in Example 10.6.3

Example 10.6.4: Finding the intersection of two planes

Give the parametric equations of the line that is the intersection of the planes p and p , where: 1 2

p1 : x − (y − 2) + (z − 1) = 0

p2 : −2(x − 2) + (y + 1) + (z − 3) = 0

Solution
To find an equation of a line, we need a point on the line and the direction of the line.
We can find a point on the line by solving each equation of the planes for z :
p1 : z = −x + y − 1

p2 : z = 2x − y − 2

We can now set these two equations equal to each other (i.e., we are finding values of x and y where the planes have the same
z value):

−x + y − 1 = 2x − y − 2

2y = 3x − 1

1
y = (3x − 1)
2

We can choose any value for x; we choose x = 1 . This determines that y = 1 . We can now use the equations of either plane to
find z : when x = 1 and y = 1 , z = −1 on both planes. We have found a point P on the line: P = (1, 1, −1).
We now need the direction of the line. Since the line lies in each plane, its direction is orthogonal to a normal vector for each
plane. Considering the equations for p and p , we can quickly determine their normal vectors. For p , n⃗ = ⟨1, −1, 1⟩ and for
1 2 1 1

p2 , n⃗
2 = ⟨−2, 1, 1⟩. A direction orthogonal to both of these directions is their cross product: d ⃗ = n⃗ 1 × n⃗ 2 = ⟨−2, −3, −1⟩.

The parametric equations of the line through P = (1, 1, −1) in the direction of d = ⟨−2, −3, −1⟩ is:
ℓ : x = −2t + 1 y = −3t + 1 z = −t − 1. (10.6.10)

The planes and line are graphed in Figure 10:58.

10.6.4 https://math.libretexts.org/@go/page/4218
Figure 10.58: Graphing the planes and their line of intersection in Example 10.6.4

Example 10.6.5: Finding the intersection of a plane and a line

Find the point of intersection, if any, of the line ℓ(t) = ⟨3, −3, −1⟩ + t⟨−1, 2, 1⟩ and the plane with equation in general form
2x + y + z = 4 .

Solution
The equation of the plane shows that the vector n⃗ = ⟨2, 1, 1⟩ is a normal vector to the plane, and the equation of the line shows
that the line moves parallel to d ⃗ = ⟨−1, 2, 1⟩ . Since these are not orthogonal, we know there is a point of intersection. (If there
were orthogonal, it would mean that the plane and line were parallel to each other, either never intersecting or the line was in
the plane itself.)
To find the point of intersection, we need to find a t value such that ℓ(t) satisfies the equation of the plane. Rewriting the
equation of the line with parametric equations will help:

⎧x = 3 −t

ℓ(t) = ⎨ y = −3 + 2t .

z = −1 + t

Replacing x, y and z in the equation of the plane with the expressions containing t found in the equation of the line allows us
to determine a t value that indicates the point of intersection:

2x + y + z = 4

2(3 − t) + (−3 + 2t) + (−1 + t) =4

t = 2.

When t = 2 , the point on the line satisfies the equation of the plane; that point is ℓ(2) = ⟨1, 1, 1⟩ . Thus the point (1, 1, 1) is
the point of intersection between the plane and the line, illustrated in Figure 10.59.

Figure 10.59: Illustrating the intersection of a line and a plane in Example 10.6.5

Distances
Just as it was useful to find distances between points and lines in the previous section, it is also often necessary to find the distance
from a point to a plane.

10.6.5 https://math.libretexts.org/@go/page/4218
Consider Figure 10.60, where a plane with normal vector n⃗ is sketched containing a point P and a point Q , not on the plane, is

given. We measure the distance from Q to the plane by measuring the length of the projection of P Q onto n⃗ . That is, we want:
→ →
→ n⃗ ⋅ P Q | n⃗ ⋅ P Q|
∥ proj P Q∥ = ∥ n⃗ ∥ = (10.6.11)
n⃗ 2
∥n∥ ∥n∥

Equation 10.6.11 is important as it does more than just give the distance between a point and a plane. We will see how it allows us
to find several other distances as well: the distance between parallel planes and the distance from a line and a plane. Because
Equation 10.6.11 is important, we restate it as a Key Idea.

Figure 10.60: Illustrating finding the distance from a point to a plane.

KEY IDEA 51 Distance from a Point to a Plane

Let a plane with normal vector n⃗ be given, and let Q be a point. The distance h from Q to the plane is

| n⃗ ⋅ P Q|
h = , (10.6.12)
∥n∥

where P is any point in the plane.

Example 10.6.6: Distance between a point and a plane

Find the distance bewteen the point Q = (2, 1, 4) and the plane with equation 2x − 5y + 6z = 9 .
Solution
Using the equation of the plane, we find the normal vector n⃗ = ⟨2, −5, 6⟩ . To find a point on the plane, we can let x and y be
anything we choose, then let z be whatever satisfies the equation. Letting x and y be 0 seems simple; this makes z = 1.5 . Thus

we let P , and P Q = ⟨2, 1, 2.5⟩.
= ⟨0, 0, 1.5⟩

The distance h from Q to the plane is given by Key Idea 51:



| n⃗ ⋅ P Q|
h =
∥n∥

|⟨2, −5, 6⟩ ⋅ ⟨2, 1, 2.5⟩|


=
∥⟨2, −5, 6⟩∥

|14|
= −−
√65

≈ 1.74.

We can use Key Idea 51 to find other distances. Given two parallel planes, we can find the distance between these planes by letting
P be a point on one plane and Q a point on the other. If ℓ is a line parallel to a plane, we can use the Key Idea to find the distance

between them as well: again, let P be a point in the plane and let Q be any point on the line. (One can also use Key Idea 50.) The
Exercise section contains problems of these types.
These past two sections have not explored lines and planes in space as an exercise of mathematical curiosity. However, there are
many, many applications of these fundamental concepts. Complex shapes can be modeled (or, approximated) using planes. For
instance, part of the exterior of an aircraft may have a complex, yet smooth, shape, and engineers will want to know how air flows
across this piece as well as how heat might build up due to air friction. Many equations that help determine air flow and heat

10.6.6 https://math.libretexts.org/@go/page/4218
dissipation are difficult to apply to arbitrary surfaces, but simple to apply to planes. By approximating a surface with millions of
small planes one can more readily model the needed behavior.

This page titled 10.6: Planes is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et al. via
source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

10.6.7 https://math.libretexts.org/@go/page/4218
10.E: Applications of Vectors (Exercises)
10.1: Introduction to Cartesian Coordinates in Space
Terms and Concepts
1. Axes drawn in space must conform to the ________ _________ rule.
2. In the plane, the equation x = 2 defines a ________; in space, x = 2 defines a ________.
3. In the plane, the equation y = x defines a ________; in space, y = x defines a ________.
2 2

4. Which quadric surface looks like a Pringles chip?


5. Consider the hyperbola x 2
−y
2
=1 in the plane. If this hyperbola is rotated about the x-axis, what quadric surface is formed?
6. Consider the hyperbola x 2
−y
2
=1 in the plane. If this hyperbola is rotated about the y-axis, what quadric surface is formed?

Problems
7. The points A = (1, 4, 2), B = (2, 6, 3) and C = (4, 3, 1) form a triangle in space. Find the distances between each pair of
points and determine if the triangle is a right triangle.
8. The points A = (1, 1, 3), B = (3, 2, 7), C = (2, 0, 8) and D = (0, −1, 4) form a quadrilateral ABCD in space. Is this a
parallelogram?
9. Find the center and radius of the sphere defined by x 2
− 8x + y
2
+ 2y + z
2
+8 = 0 .
10. Find the center and radius of the sphere defined by x 2
+y
2
+z
2
+ 4x − 2y − 4z + 4 = 0 .
In Exercises 11-14, describe the region in space defined by the inequalities.
11. x 2
+y
2
+z
2
<1

12. 0 ≤ x ≤ 3
13. x ≥ 0, y ≥ 0, z ≥ 0
14. y ≥ 3
In Exercises 15-18, sketch the cylinder in space.
15. z = x 3

16. y = cos z
2
2
y
17. x

4
+
9
=1

18. y = 1

In Exercises 19-22, give the equation of the surface of revolution described.


19. Revolve z = 1

1+y
2
about the y-axis.

20. Revolve y = x about the x-axis.2

21. Revolve z = x about the z-axis.


2

22. Revolve z = 1/x about the z-axis.


In Exercises 23-26, a quadric surface is sketched. Determine which of the given equations best fits the graph.

10.E.1 https://math.libretexts.org/@go/page/9999
23.

24.

25.

26.
In Exercises 27-32, sketch the quadric surface.
27. z − y 2
+x
2
=0

2
y
28. z 2
=x
2
+
4

29. x = −y 2
−z
2

30. 16x 2
− 16 y
2
− 16 z
2
=1

2 2

31. x

9
−y
2
+
z

25
=1

32. 4x 2
+ 2y
2
+z
2
=4

10.2: An Introduction to Vectors

10.E.2 https://math.libretexts.org/@go/page/9999
Terms and Concepts
1. Name two different things that cannot be described with just one number, but rather need 2 or more numbers to fully describe
them.
2. What is the difference between (1, 2) and ⟨1, 2⟩?
3. What is a unit vector?
4. What does it mean for two vectors to be parallel?
5. What effect does multiplying by a vector by -2 have?

Problems

In Exercises 6-9, points P and Q are given. Write the vector P Q in component form and using the standard unit vectors.
6. P = (2, −1), Q = (3, 5)

7. P = (3, 2), Q = (7, −2)

8. P = (0, 3, −1), Q = (6, 2, 5)

9. P = (2, 1, 2), Q = (4, 3, 2)

10. Let u⃗ = ⟨1, −2⟩ and v ⃗ = ⟨1, 1⟩ .


(a) Find u⃗ + v,⃗ u⃗ − v,⃗ 2u⃗ − 3v ⃗ .
(b) Sketch the above vectors on the same axes, along with u⃗ and v ⃗ .
(c) Find x⃗ where u⃗ + x⃗ = 2v ⃗ − x⃗ .
11. Let u⃗ = ⟨1, 1, −1⟩ and v ⃗ = ⟨2, 1, 2⟩ .

(a) Find u⃗ + v,⃗ u⃗ − v,⃗ π u⃗ − √2v ⃗ .
(b) Sketch the above vectors on the same axes, along with u⃗ and v ⃗ .
(c) Find x⃗ where u⃗ + x⃗ = v ⃗ + 2x⃗ .
In Exercises 12-15, sketch u⃗, v,⃗ u⃗ + v ⃗ and u⃗ − v ⃗ on the same axes.
12.

13.

10.E.3 https://math.libretexts.org/@go/page/9999
14.

15.

In Exercises 16-19, find ∥u⃗∥, ∥v∥,


⃗ ∥ u⃗ + v∥ ⃗ .
⃗ and ∥ u⃗ − v∥

16. u⃗ = ⟨2, 1⟩, v ⃗ = ⟨3, −2⟩

17. u⃗ = ⟨−3, 2, 2⟩, v ⃗ = ⟨1, −1, 1⟩

18. u⃗ = ⟨1, 2⟩, v ⃗ = ⟨−3, −6⟩

19. u⃗ = ⟨2, −3, 6⟩, v ⃗ = ⟨10, −15, 30⟩

20. Under what conditions is ∥u⃗∥ + ∥v∥⃗ = ∥u⃗ + v∥⃗ ?


In Exercises 21-24, find the unit vector u⃗ in the direction of v ⃗.
21. v ⃗ = ⟨3, 7⟩
22. v ⃗ = ⟨6, 8⟩
23. v ⃗ = ⟨1, −2, 2⟩
24. v ⃗ = ⟨2, −2, 2⟩
25. Find the unit vector in the first quadrant of R that makes a 50 angle with the x-axis.
2 ∘

26. Find the unit vector in the second quadrant of R that makes a 30 angle with the y-axis.
2 ∘

27. Verify, from Key Idea 48, that u⃗ = ⟨sin θ cos ϕ, sin θ sin ϕ, cos θ⟩ is a unit vector for all angles θ and ϕ .
A weight of 100lb is suspended from two chains, making an angles with the vertical of θ and ϕ as shown in the figure below.

In Exercises 28-31, angles θ and ϕ are given. Find the force applied to each chain.
28. θ = 30 ∘
, ϕ = 30

29. θ = 60 ∘
, ϕ = 60

30. θ = 20 ∘
, ϕ = 15

31. θ = 0 ∘
, ϕ =0

A weight of plb is suspended from a chain of length l while a constant force of F ⃗ pushes the weight to the right, making an
w

angle of θ with the vertical, as shown in the figure below.

10.E.4 https://math.libretexts.org/@go/page/9999
In Exercises 32-35, a force F ⃗ and length l are given. Find the angle θ and the height the weight is lifted as it moves to the
w

right.
32. F ⃗
w = 1lb, l = 1ft, p = 1lb

33. F ⃗
w = 1lb, l = 1ft, p = 10lb

34. F ⃗
w = 1lb, l = 10ft, p = 1lb

35. F ⃗
w = 10lb, l = 10ft, p = 1lb

10.3: The Dot Product


Terms and Concepts
1. The dot product of two vectors is a __________, not a vector.
2. How are the concepts of the dot product and vector magnitude related?
3. How can one quickly tell if the angle between two vectors is acute or obtuse?
4. Give a synonym for "orthogonal."

Problems
In Exercises 5-10, find the dot product of the given vectors.
5. u⃗ = ⟨2, −4⟩, v⃗ = ⟨3, 7⟩
6. u⃗ = ⟨5, 3⟩, v ⃗ = ⟨6, 1⟩
7. u⃗ = ⟨1, −1, 2⟩, v ⃗ = ⟨2, 5, 3⟩
8. u⃗ = ⟨3, 5, −1⟩, v ⃗ = ⟨4, −1, 7⟩
9. u⃗ = ⟨1, 1⟩, v ⃗ = ⟨1, 2, 3⟩
10. u⃗ = ⟨1, 2, 3⟩, v ⃗ = ⟨0, 0, 0⟩
11. Create your own vectors u⃗, v,⃗ and w⃗ in R
2
and show that u⃗ ⋅ (v ⃗ + w⃗) = u⃗ ⋅ v ⃗ + u⃗ ⋅ w⃗ .
12. Create your own vectors u⃗ and v ⃗ in R and scalar c and show that c(u⃗ ⋅ v)⃗ = u⃗ ⋅ (cv)⃗ .
3

In Exercises 13-16, find the measure of the angle between the two vectors in both radians and degrees.
13. u⃗ = ⟨1, 1⟩, v ⃗ = ⟨1, 2⟩

14. u⃗ = ⟨−2, 1⟩, v ⃗ = ⟨3, 5⟩

15. u⃗ = ⟨8, 1, −4⟩, v ⃗ = ⟨2, 2, 0⟩

16. u⃗ = ⟨1, 7, 2⟩, v ⃗ = ⟨4, −2, 5⟩

In Exercises 17-20, a vectors v ⃗ is given. Give two vectors that are orthogonal to v ⃗.
17. v ⃗ = ⟨4, 7⟩
18. v ⃗ = ⟨−3, 5⟩
19. v ⃗ = ⟨1, 1, 1⟩
20. v ⃗ = ⟨1, −2, 3⟩

10.E.5 https://math.libretexts.org/@go/page/9999
In Exercises 21-26, vectors u⃗ and v ⃗ are given. Find , the orthogonal projection of
proj v ⃗ u⃗ u⃗ onto v ⃗, and sketch all three
vectors on the same axes.
21. u⃗ = ⟨1, 2⟩, v⃗ = ⟨−1, 3⟩
22. u⃗ = ⟨5, 5⟩, v ⃗ = ⟨1, 3⟩
23. u⃗ = ⟨−3, 2⟩, v ⃗ = ⟨1, 1⟩
24. u⃗ = ⟨−3, 2⟩, v ⃗ = ⟨2, 3⟩
25. u⃗ = ⟨1, 5, 1⟩, v ⃗ = ⟨1, 2, 3⟩
26. u⃗ = ⟨3, −1, 2⟩, v ⃗ = ⟨2, 2, 1⟩
In Exercises 27-32, vectors u⃗ and vecv are given. Write u⃗ as the sum of two vectors, one of which is parallel to v ⃗ and one of
which is perpendicular to v ⃗. Note: these are the same pairs of vectors as found in Exercises 21-26.
27. u⃗ = ⟨1, 2⟩, v ⃗ = ⟨−1, 3⟩
28. u⃗ = ⟨5, 5⟩, v ⃗ = ⟨1, 3⟩
29. u⃗ = ⟨−3, 2⟩, v ⃗ = ⟨1, 1⟩
30. u⃗ = ⟨−3, 2⟩, v ⃗ = ⟨2, 3⟩
31. u⃗ = ⟨1, 5, 1⟩, v ⃗ = ⟨1, 2, 3⟩
32. u⃗ = ⟨3, −1, 2⟩, v ⃗ = ⟨2, 2, 1⟩
33. A 10lb box sits on a ramp that rises 4ft over a distance of 20ft. How much force is required to keep the box from sliding down
the ramp?
34. A 10lb box sits on a 15ft ramp that makes a 30 angle with the horizontal. How much force is required to keep the box from

sliding down the ramp?


35. How much work is performed in moving a box horizontally 10ft with a force of 20lb applied at an angle of 45 to the ∘

horizontal?
36. How much work is performed in moving a box horizontally 10ft with a force of 20lb applied at an angle of 10 to the ∘

horizontal?
37. How much work is performed in moving a box up the length of a ramp that rises 2ft over a distance of 10ft, with a force of 50lb
applied horizontally?
38. How much work is performed in moving a box up the length of a ramp that rises 2ft over a distance of 10ft, with a force of 50lb
applied at an angle of 45 to the horizontal?

39. How much work is performed in moving a box up the length of a 10ft ramp that makes a 5 angle with the horizontal, with 50lb

of force applied in the direction of the ramp?

10.4: The Cross Product


Terms and Concepts
1. The cross product of two vectors is a ________, not a scalar.
2. One can visualize the direction of u⃗ × v ⃗ using the _______ ________ ________.
3. Give synonym for "orthogonal."
4. T/F: A fundamental principle of the cross product is that u⃗ × v⃗ is orthogonal to u⃗ and v⃗.
5. _______ is a measure of the turning force applied to an object.

Problems
In Exercises 6-14, vectors u⃗ and v ⃗ are given. Compute u⃗ × v ⃗ and show this is orthogonal to both u⃗ and v ⃗.
6. u⃗ = ⟨3, 2, −2⟩, v ⃗ = ⟨0, 1, 5⟩

10.E.6 https://math.libretexts.org/@go/page/9999
7. u⃗ = ⟨5, −4, 3⟩, v ⃗ = ⟨2, −5, 1⟩

8. u⃗ = ⟨4, −5, −5⟩, v⃗ = ⟨3, 3, 4⟩

9. u⃗ = ⟨−4, 7, −10⟩, v ⃗ = ⟨4, 4, 1⟩

10. u⃗ = ⟨1, 0, 1⟩, v ⃗ = ⟨5, 0, 7⟩

11. u⃗ = ⟨1, 5, −4⟩, v ⃗ = ⟨−2, −10, 8⟩

12. u⃗ = i ,⃗ v⃗ = j

13. u⃗ = i ,⃗ v⃗ = k

14. u⃗ = j,⃗ v⃗ = j

15. Pick any vectors u⃗, v ⃗ and w⃗ in R and show that u⃗ × (v ⃗ + w⃗ = u⃗ × v ⃗ + u⃗ × w⃗ .


3

16. Pick any vectors u⃗, v ⃗ and w⃗ in R and show that u⃗ ⋅ (v ⃗ × w⃗ = (u⃗ × v)⃗ ⋅ w⃗ .
3

In Exercises 17-20, the magnitude of vectors u⃗ and v⃗ in R


3
are given, along with the angle θ between them. Use this
information to find the magnitude of u⃗ × v ⃗ .
17. ∥u⃗∥ = 2, ⃗ = 5,
∥ v∥ θ = 30

18. ∥u⃗∥ = 3, ⃗ = 7,
∥ v∥ θ = π/2

19. ∥u⃗∥ = 3, ⃗ = 4,
∥ v∥ θ =π

20. ∥u⃗∥ = 2, ⃗ = 5,
∥ v∥ θ = 5π/6

In Exercises 21-24, find the area of the parallelogram defined by the given vectors.
21. u⃗ = ⟨1, 1, 2⟩, v ⃗ = ⟨2, 0, 3⟩

22. u⃗ = ⟨−2, 1, 5⟩, v ⃗ = ⟨−1, 3, 1⟩

23. u⃗ = ⟨1, 2⟩, v ⃗ = ⟨2, 1⟩

24. u⃗ = ⟨2, 0⟩, v ⃗ = ⟨0, 3⟩

In Exercises 25-28, find the area of the triangle with the given vertices.
25. Vertices: (0, 0, 0), (1, 3, −1) and (2, 1, 1).
26. Vertices: (5, 2, −1), (3, 6, 2) and (1, 0, 4).
27. Vertices: (1, 1), (1, 3) and (2, 2).
28. Vertices: (3, 1), (1, 2) and (4, 3).
In Exercises 29-30, find the area of the quadrilateral with the given vertices. (Hint: break the quadrilateral into the 2
triangles.)
29. Vertices: (0, 0), (1, 2), (3, 0) and (4, 3).
30. Vertices: (0, 0, 0), (2, 1, 1), (−1, 2, −8) and (1, −1, 5).
In Exercises 31-32, find the volume of the parallelepiped defined by the given vectors.
31. u⃗ = ⟨1, 1, 1⟩, v ⃗ = ⟨1, 2, 3⟩, w⃗ = ⟨1, 0, 1⟩

32. u⃗ = ⟨−1, 2, 1⟩, v⃗ = ⟨2, 2, 1⟩, w⃗ = ⟨3, 1, 3⟩

In Exercises 33-36, find a unit vector orthogonal to both u⃗ and v ⃗.


33. u⃗ = ⟨1, 1, 1⟩, v ⃗ = ⟨2, 0, 1⟩

34. u⃗ = ⟨1, −2, 1⟩, v ⃗ = ⟨3, 2, 1⟩

35. u⃗ = ⟨5, 0, 2⟩, v ⃗ = ⟨−3, 0, 7⟩

36. u⃗ = ⟨1, −2, 1⟩, v ⃗ = ⟨−2, 4, −2⟩

10.E.7 https://math.libretexts.org/@go/page/9999
37. A bicycle rider applies 150lb of force, straight down, onto a pedal that extends ϳin horizontally from the crankshaft. Find the
magnitude of the torque applied to the crankshaft.
38. A bicycle rider applies 150lb of force, straight down, onto a pedal that extends 7in from the crankshaft, making a 30 angle ∘

with the horizontal. Find the magnitude of the torque applied to the crankshaft.
39. To turn a stubborn bolt, 80lb of force is applied to a 10in wrench. What is the maximum amount of torque that can be applied to
the bolt?
40. To turn a stubborn bolt, 80lb of force is applied to a 10in wrench in a confined space, where the direction of applied force
makes a 10 angle with the wrench. How much torque is subsequently applied to the wrench?

41. Show, using the definition of the Cross Product, that u⃗ ⋅ (u⃗ × v)⃗ = 0 ; that is, that u⃗ is orthogonal to the cross product of u⃗ and
v ⃗.

42. Show, using the definition of the Cross Product, that vecu × u⃗ = 0⃗ .

10.5: Lines
Terms and Concepts
1. To find an equation of a line, what two pieces of information are needed?
2. Two distinct lines in the plane can intersect or be _______.
3. Two distinct lines in space can intersect, be _________ or be __________.
4. Use your own words to describe what it means for two lines in space to be skew.

Problems
In Exercises 5-14, write the vector, parametric and symmetric equations of the lines described.
5. Passes through P = (2, −4, 1) , parallel to d ⃗ = ⟨9, 2, 5⟩ .

6. Passes through P = (6, 1, 7) , parallel to d ⃗ = ⟨−3, 2, 5⟩ .


7. Passes through P = (2, 1, 5) and Q = (7, −2, 4).
8. Passes through P = (1, −2, 3) and Q = (5, 5, 5).

9. Passes through P = (0, 1, 2) and orthogonal to both d ⃗ 1 = ⟨2, −1, 7⟩ and d ⃗ 2 = ⟨7, 1, 3⟩ .

10. Passes through P = (5, 1, 9) and orthogonal to both d ⃗ 1 = ⟨1, 0, 1⟩ and d ⃗


2 = ⟨2, 0, 3⟩ .

11. Passes through the point of intersection of l ⃗ 1 (t) and l ⃗ 2 (t) and orthogonal to both lines, where
l (t) = ⟨2, 1, 1⟩ + t⟨5, 1, −2⟩ and

1

l (t) = ⟨−2, −1, 2⟩ + t⟨3, 1, −1⟩ .



2

12. Passes through the point of intersection of l ⃗ 1 (t) and l



2 (t) and orthogonal to both lines, where
⎧x =t ⎧ x = 2 +t

l1 = ⎨ y = −2 + 2t and l2 = ⎨ y = 2 − t
⎩ ⎩
z = 1 +t z = 1 +t

13. Passes through P = (1, 1) , parallel to d ⃗ = ⟨2, 3⟩ .

14. Passes through P = (−2, 5) , parallel to d ⃗ = ⟨0, 1⟩ .


In Exercises 15-22, determine if the described lines are the same line, parallel lines, intersecting or skew lines. If
intersecting, give the point of intersection.
15.
l

1 (t) = ⟨1, 2, 1⟩ + t⟨2, −1, 1⟩, l

2 (t) = ⟨3, 3, 3⟩ + t⟨−4, 2, −2⟩ .
16.
l

1 (t) = ⟨2, 1, 1⟩ + t⟨5, 1, 3⟩, l

2 (t) = ⟨14, 5, 9⟩ + t⟨1, 1, 1⟩ .

10.E.8 https://math.libretexts.org/@go/page/9999
17.
l

1 (t) = ⟨3, 4, 1⟩ + t⟨2, −3, 4⟩, l

2 (t) = ⟨−3, 3, −3⟩ + t⟨3, −2, 4⟩ .
18.
l

1 (t) = ⟨1, 1, 1⟩ + t⟨3, 1, 3⟩, l

2 (t) = ⟨7, 3, 7⟩ + t⟨6, 2, 6⟩ .
19.
⎧ x = 1 + 2t ⎧ x = 3 −t

l1 = ⎨ y = 3 − 2t and l2 = ⎨ y = 2 + 5t
⎩ ⎩
z =t z = 2 + 7t

20.
⎧ x = 1.1 + 0.6t ⎧ x = 3.11 + 3.4t

l1 = ⎨ y = 3.77 + 0.9t and l2 = ⎨ y = 2 + 5.1t


⎩ ⎩
z = −2.3 + 1.5t z = 2.5 + 8.5t

21.
⎧ x = −0.2 + 0.6t ⎧ x = 0.86 + 9.2t

l1 = ⎨ y = 1.33 − 0.45t and l2 = ⎨ y = 0.835 + 9.2t


⎩ ⎩
z = −4.2 + 1.05t z = −3.045 + 16.1t

22.
⎧ x = 0.1 + 1.1t ⎧ x = 4 − 2.1t

l1 = ⎨ y = 2.9 − 1.5t and l2 = ⎨ y = 1.8 + 7.2t


⎩ ⎩
z = 3.2 + 1.6t z = 3.1 + 1.1t

In Exercises 23-26, find the distance from the point to the line.

23. P = (1, 1, 1),



l (t) = ⟨2, 1, 3⟩ + t⟨2, 1, −2⟩

24. P = (2, 5, 6),



l (t) = ⟨−1, 1, 1⟩ + t⟨1, 0, 1⟩

25. P = (0, 3),



l (t) = ⟨2, 0⟩ + t⟨1, 1⟩

26. P = (1, 1),



l (t) = ⟨4, 5⟩ + t⟨−4, 3⟩

In Exercises 27-28, find the distance between the two lines.


27.
l

1 (t) = ⟨1, 2, 1⟩ + t⟨2, −1, 1⟩, l

2 (t) = ⟨3, 3, 3⟩ + t⟨4, 2, −2⟩ .
28.
l

1 (t) = ⟨0, 0, 1⟩ + t⟨1, 0, 0⟩, l

2 (t) = ⟨0, 0, 3⟩ + t⟨0, 1, 0⟩ .
Exercises 29-31 explore special cases of the distance formulas found in Key Idea 50.
29. Let Q be a point on the line l(t). Show why the distance formula correctly gives the distance from the point to the line as 0.
30. Let lines l 1 (t) and l2 (t) be intersecting lines. Show why the distance formula correctly gives the distance between these lines
as 0.
31. Let lines l (t) and l (t) be parallel.
1 2

(a) Show why the distance formula for distance between lines cannot be used as stated to find the distance between the lines.
→ → →
(b) Show why letting c = (P P × d ) × d allows one to the use the formula.
1 2 2 2

(c) Show how one can use the formula for the distance between a point and a line and a line to find the distance between parallel
lines.

10.6: Planes
Terms and Concepts
1. In order to find the equation of a plane, what two pieces of information must one have?
2. What is the relationship between a plane and one of its normal vectors?

10.E.9 https://math.libretexts.org/@go/page/9999
Problems
In Exercises 3-6, give any two points in the given plane.
3. 2x − 4y + 7z = 2
4. 3(x + 2) + 5(y − 9) − 4z = 0
5. x = 2
6. 4(y + 2) − (z − 6) = 0
In Exercises 7-20, give the equation of the described plane in standard and general forms.
7. Passes through (2, 3, 4) and has normal vector n⃗ = ⟨3, −1, 7⟩ .
8. Passes through (1, 3, 5) and has normal vector n⃗ = ⟨0, 2, 4⟩.
9. Passes through the points (1, 2, 3), (3, −1, 4) and (1, 0, 1).
10. Passes through the points (5, 3, 8), (6, 4, 9) and (3, 3, 3).
11. Contains the intersecting lines
l (t) = ⟨2, 1, 2⟩ + t⟨1, 2, 3⟩ and
1

l (t) = ⟨2, 1, 2⟩ + t⟨2, 5, 4⟩ .


2

12. Contains the intersecting lines


l (t) = ⟨5, 0, 3⟩ + t⟨−1, 1, 1⟩ and
1

l (t) = ⟨1, 4, 7⟩ + t⟨3, 0, −3⟩ .


2

13. Contains the parallel lines


l (t) = ⟨1, 1, 1⟩ + t⟨1, 2, 3⟩ and
1

l (t) = ⟨1, 1, 2⟩ + t⟨1, 2, 3⟩ .


2

14. Contains the parallel lines


l (t) = ⟨2, 1, 2⟩ + t⟨1, 2, 3⟩ and
1

l (t) = ⟨2, 1, 2⟩ + t⟨2, 5, 4⟩ .


2

15. Contains the point (2, −6, 1) and the line


⎧ x = 2 + 5t

l(t) = ⎨ y = 2 + 2t

z = −1 + 2t

16. Contains the point (5, 7, 3) and the line


⎧x =t

l(t) = ⎨ y = t

z =t

17. Contains the point (5, 7, 3) and is orthogonal to the line l(t) = ⟨4, 5, 6⟩ + t⟨1, 1, 1⟩ .
18. Contains the point (4, 1, 1) and is orthogonal to the line
⎧ x = 4 + 4t

l(t) = ⎨ y = 1 + 1t

z = 1 + 1t

19. Contains the point (−4, 7, 2) and is parallel to the plane 3(x − 2) + 8(y + 1) − 10z = 0 .
20. Contains the point (1, 2, 3) and is parallel to the plane x = 5 .
In Exercises 21-22, give the equation of the line that is the intersection of the given plane.
21.
p1 : 3(x − 2) + (y − 1) + 4z = 0, and

p2 : 2(x − 1) − 2(y + 3) + 6(z − 1) = 0 .


22.
p1 : 5(x − 5) + 2(y + 2) + 4(z − 1) = 0, and

p2 : 3x − 4(y − 1) + 2(z − 1) = 0 .

10.E.10 https://math.libretexts.org/@go/page/9999
In Exercises 23-26, find the point of intersection between the line and the plane.
23.
line: ⟨5, 1, −1⟩ + t⟨2, 2, 1⟩,
plane: 5x − y − z = −3
24.
line: ⟨4, 1, 0⟩ + t⟨1, 0, −1⟩,
plane: 3x + y − 2x = 8
25.
line: ⟨1, 2, 3⟩ + t⟨3, 5, −1⟩,
plane: 3x − 2y − z = 4
26.
line: ⟨1, 2, 3⟩ + t⟨3, 5, −1⟩,
plane: 3x − 2y − z = −4
In Exercises 27-30, find the given distances.
27. The distance from the point (1, 2, 3) to the plane
3(x − 1) + (y − 2) + 5(z − 2) = 0 .
28. The distance from the point (2, 6, 2) to the plane
2(x − 1) − y + 4(z + 1) = 0 .

29. The distance between the parallel planes


x + y + z = 0 and

(x − 2) + (y − 3) + (z + 4) = 0 .
30. The distance between parallel planes
2(x − 1) + 2(y + 1) + (z − 2) = 0 and

2(x − 3) + 2(y − 1) + (z − 3) = 0

31. Show why if the point Q lies in a plane, then the distance formula correctly gives the distance from the point to the plane as 0.
32. How is Exercise 30 in Section 10.5 easier to answer once we have an understanding of planes?

10.E: Applications of Vectors (Exercises) is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by LibreTexts.

10.E.11 https://math.libretexts.org/@go/page/9999
CHAPTER OVERVIEW
11: Vector-Valued Functions
In the previous chapter, we learned about vectors and were introduced to the power of vectors within mathematics. In this chapter,
we’ll build on this foundation to define functions whose input is a real number and whose output is a vector. We’ll see how to
graph these functions and apply calculus techniques to analyze their behavior. Most importantly, we’ll see why we are interested in
doing this: we’ll see beautiful applications to the study of moving objects.
11.1: Vector–Valued Functions
11.2: Calculus and Vector-Valued Functions
11.3: The Calculus of Motion
11.4: Unit Tangent and Normal Vectors
11.5: The Arc Length Parameter and Curvature
11.E: Applications of Vector Valued Functions (Exercises)

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 11: Vector-Valued Functions is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory
Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

1
11.1: Vector–Valued Functions
We are very familiar with real valued functions, that is, functions whose output is a real number. This section introduces vector-
valued functions - functions whose output is a vector.

Definition 11.1.1: Vector-Valued Functions

A vector-valued function is a function of the form



r (t) = ⟨ f (t), g(t) ⟩ (11.1.1)

or

r (t) = ⟨ f (t), g(t), h(t) ⟩, (11.1.2)

where f , g and h are real valued functions.


The domain of ⇀
r is the set of all values of t for which ⇀
r (t) is defined. The range of ⇀
r is the set of all possible output vectors
r (t) .

Evaluating and Graphing Vector-Valued Functions


Evaluating a vector-valued function at a specific value of t is straightforward; simply evaluate each component function at that
value of t . For instance, if r (t) = ⟨t , t + t − 1⟩ , then r (−2) = ⟨4, 1⟩ . We can sketch this vector, as is done in Figure 11.1.1a.
⇀ 2 2 ⇀

Plotting lots of vectors is cumbersome, though, so generally we do not sketch the whole vector but just the terminal point. The
graph of a vector-valued function is the set of all terminal points of r (t) , where the initial point of each vector is always the

origin. In Figure 11.1.1b we sketch the graph of r ; we can indicate individual points on the graph with their respective vector, as

shown.

Figure 11.1.1 : Sketching the graph of a vector-valued function.


Vector-valued functions are closely related to parametric equations of graphs. While in both methods we plot points (x(t), y(t)) or
(x(t), y(t), z(t)) to produce a graph, in the context of vector-valued functions each such point represents a vector. The

implications of this will be more fully realized in the next section as we apply calculus ideas to these functions.

Example 11.1.1: Graphing vector-valued functions


1
Graph ⇀ 3
r (t) = ⟨t − t,
2
⟩ , for −2 ≤ t ≤ 2 . Sketch ⇀
r (−1) and ⇀
r (2) .
t +1

Solution
We start by making a table of t , x and y values as shown in Figure 11.1.1a. Plotting these points gives an indication of what
the graph looks like. In Figure 11.1.1b, we indicate these points and sketch the full graph. We also highlight r (−1) and r (2) ⇀ ⇀

on the graph.

11.1.1 https://math.libretexts.org/@go/page/4220
Figure 11.1.2 : Sketching the vector-valued function of Example 11.1.1

Example 11.1.2: Graphing vector-valued functions.

Graph ⇀
r (t) = ⟨cos t, sin t, t⟩ for 0 ≤ t ≤ 4π .

Solution
We can again plot points, but careful consideration of this function is very revealing. Momentarily ignoring the third
component, we see the x and y components trace out a circle of radius 1 centered at the origin. Noticing that the z component
is t , we see that as the graph winds around the z -axis, it is also increasing at a constant rate in the positive z direction, forming
a spiral. This is graphed in Figure 11.1.3. In the graph r (7π/4) ≈ (0.707, −0.707, 5.498)is highlighted to help us understand

the graph.

Figure 11.1.3 : Viewing a vector-valued function, and its derivative at one point.

Algebra of Vector-Valued Functions

Definition 11.1.2: Operations on Vector-Valued Functions

Let ⇀
r 1 (t) = ⟨f1 (t), g1 (t)⟩ and ⇀
r 2 (t) = ⟨f2 (t), g2 (t)⟩ be vector-valued functions in R and let c be a scalar. Then:
2

1. r (t) ± r (t) = ⟨ f (t) ± f



1

2 1 2 (t), g1 (t) ± g2 (t) ⟩ .
2. c r (t) = ⟨ cf (t), cg (t) ⟩ .

1 1 1

11.1.2 https://math.libretexts.org/@go/page/4220
A similar definition holds for vector-valued functions in R . 3

This definition states that we add, subtract and scale vector-valued functions component-wise. Combining vector-valued functions
in this way can be very useful (as well as create interesting graphs).

Example 11.1.3: Adding and scaling vector-valued functions.

Let ⇀
r 1 (t) = ⟨ 0.2t, 0.3t ⟩ , ⇀
r 2 (t) = ⟨ cos t, sin t ⟩ and ⇀ ⇀ ⇀
r (t) = r 1 (t) + r 2 (t) . Graph ⇀
r 1 (t) , ⇀
r 2 (t) , ⇀
r (t) and ⇀
5 r (t) on
−10 ≤ t ≤ 10 .
Solution
We can graph r and r easily by plotting points (or just using technology). Let's think about each for a moment to better

1

2

understand how vector-valued functions work.


We can rewrite r (t) = ⟨ 0.2t, 0.3t ⟩ as r (t) = t⟨0.2, 0.3⟩. That is, the function

1

1

r1 scales the vector ⟨0.2, 0.3⟩ by t . This
scaling of a vector produces a line in the direction of ⟨0.2, 0.3⟩.
We are familiar with ⇀
r 2 (t) = ⟨ cos t, sin t ⟩ ; it traces out a circle, centered at the origin, of radius 1. Figure 11.1.4a graphs
r (t) and r (t) .
⇀ ⇀
1 2

Adding r (t) to r (t) produces r (t) = ⟨ cos t + 0.2t, sin t + 0.3t ⟩ , graphed in Figure 11.1.4b. The linear movement of the

1

2

line combines with the circle to create loops that move in the direction of ⟨0.2, 0.3⟩. (We encourage the reader to experiment
by changing r (t) to ⟨2t, 3t⟩, etc., and observe the effects on the loops.)

1

Figure 11.1.4 : Graphing the functions in Example 11.1.3


Multiplying r (t) by 5 scales the function by 5, producing 5 r (t) = ⟨5 cos t + 1, 5 sin t + 1.5⟩ , which is graphed in Figure
⇀ ⇀

11.1.4c along with r (t) . The new function is "5 times bigger'' than r (t) . Note how the graph of 5 r (t) in (c) looks identical to
⇀ ⇀ ⇀

the graph of r (t) in (b). This is due to the fact that the x and y bounds of the plot in (c) are exactly 5 times larger than the

bounds in (b).

Example 11.1.4: Adding and scaling vector-valued functions.

A cycloid is a graph traced by a point p on a rolling circle, as shown in Figure 11.1.5. Find an equation describing the cycloid,
where the circle has radius 1.

Figure 11.1.5 : Tracing a cycloid.


Solution
This problem is not very difficult if we approach it in a clever way. We start by letting p (t) describe the position of the point p

on the circle, where the circle is centered at the origin and only rotates clockwise (i.e., it does not roll). This is relatively simple
given our previous experiences with parametric equations; p (t) = ⟨cos t, − sin t⟩ .

We now want the circle to roll. We represent this by letting c (t) represent the location of the center of the circle. It should be

clear that the y component of c (t) should be 1; the center of the circle is always going to be 1 if it rolls on a horizontal

surface.

11.1.3 https://math.libretexts.org/@go/page/4220
The x component of c (t) is a linear function of t : f (t) = mt for some scalar m. When t = 0 , f (t) = 0 (the circle starts

centered on the y -axis). When t = 2π , the circle has made one complete revolution, traveling a distance equal to its
circumference, which is also 2π. This gives us a point on our line f (t) = mt , the point (2π, 2π). It should be clear that m = 1
and f (t) = t . So c (t) = ⟨t, 1⟩ .

We now combine p and ⇀ ⇀


c together to form the equation of the cycloid:
⇀ ⇀ ⇀
r (t) = p (t) + c (t) = ⟨cos t + t, − sin t + 1⟩,

which is graphed in Figure 11.1.6.

Figure 11.1.6 : The cycloid in Example 11.1.4 .

Displacement
A vector-valued function r (t) is often used to describe the position of a moving object at time t . At t = t , the object is at r (t );

0

0

at t = t , the object is at r (t ). Knowing the locations r (t ) and r (t ) give no indication of the path taken between them, but
1

1

0

1

often we only care about the difference of the locations, r (t ) − r (t ) , the displacement.

1

0

Definition 11.1.3: Displacement



Let ⇀
r (t) be a vector-valued function and let t 0 < t1 be values in the domain. The displacement d of ⇀
r , from t = t to t = t ,
0 1

is

⇀ ⇀
d = r (t1 ) − r (t0 ). (11.1.3)

When the displacement vector is drawn with initial point at ⇀


r (t0 ) , its terminal point is ⇀
. We think of it as the vector which
r (t1 )

points from a starting position to an ending position.

Example 11.1.5: Finding and graphing displacement vectors


π π
Let ⇀
r (t) = ⟨cos( t), sin( t)⟩ . Graph ⇀
r (t) on −1 ≤ t ≤ 1 , and find the displacement of ⇀
r (t) on this interval.
2 2

Solution
The function r (t) traces out the unit circle, though at a different rate than the "usual'' ⟨cos t, sin t⟩ parametrization. At

t = −1 , we have r (t ) = ⟨0, −1⟩ ; at t = 1 , we have r (t ) = ⟨0, 1⟩ . The displacement of r (t) on [−1, 1] is thus
⇀ ⇀ ⇀
0 0 1 1


d = ⟨0, 1⟩ − ⟨0, −1⟩ = ⟨0, 2⟩.

Figure 11.1.7 : Graphing the displacement of a position function in Example 11.1.5 .

11.1.4 https://math.libretexts.org/@go/page/4220

A graph of ⇀
r (t) on [−1, 1] is given in Figure 11.1.7, along with the displacement vector d on this interval.

Measuring displacement makes us contemplate related, yet very different, concepts. Considering the semi-circular path the object
in Example 11.1.5 took, we can quickly verify that the object ended up a distance of 2 units from its initial location. That is, we can
compute ∥d∥ = 2 . However, measuring distance from the starting point is different from measuring distance traveled. Being a
semi-circle, we can measure the distance traveled by this object as π ≈ 3.14 units. Knowing distance from the starting point allows
us to compute average rate of change.

Definition 11.1.4: Average Rate of Change

Let r (t) be a vector-valued function, where each of its component functions is continuous on its domain, and let t

0 < t1 . The
average rate of change of r (t) on [t , t ] is

0 1

⇀ ⇀
r (t1 ) − r (t0 )
average rate of change = . (11.1.4)
t1 − t0

Example 11.1.6: Average rate of change


π π
Let ⇀
r (t) = ⟨cos( t), sin( t)⟩ as in Example 11.1.5. Find the average rate of change of ⇀
r (t) on [−1, 1] and on [−1, 5].
2 2

Solution

We computed in Example 11.1.5 that the displacement of ⇀
r (t) on [−1, 1] was d = ⟨0, 2⟩ . Thus the average rate of change of
r (t) on [−1, 1] is:

⇀ ⇀
r (1) − r (−1) ⟨0, 2⟩
= = ⟨0, 1⟩.
1 − (−1) 2

We interpret this as follows: the object followed a semi-circular path, meaning it moved towards the right then moved back to
the left, while climbing slowly, then quickly, then slowly again. On average, however, it progressed straight up at a constant
rate of ⟨0, 1⟩ per unit of time.

We can quickly see that the displacement on [−1, 5] is the same as on [−1, 1], so d = ⟨0, 2⟩ . The average rate of change is
different, though:
⇀ ⇀
r (5) − r (−1) ⟨0, 2⟩
= = ⟨0, 1/3⟩.
5 − (−1) 6

As it took "3 times as long'' to arrive at the same place, this average rate of change on [−1, 5] is 1/3 the average rate of change
on [−1, 1].

We considered average rates of change in Sections 1.1 and 2.1 as we studied limits and derivatives. The same is true here; in the
following section we apply calculus concepts to vector-valued functions as we find limits, derivatives, and integrals. Understanding
the average rate of change will give us an understanding of the derivative; displacement gives us one application of integration.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 11.1: Vector–Valued Functions is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory
Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

11.1.5 https://math.libretexts.org/@go/page/4220
11.2: Calculus and Vector-Valued Functions
The previous section introduced us to a new mathematical object, the vector-valued function. We now apply calculus concepts to
these functions. We start with the limit, then work our way through derivatives to integrals.

Limits of Vector-Valued Functions


The initial definition of the limit of a vector-valued function is a bit intimidating, as was the definition of the limit in Definition 1.
The theorem following the definition shows that in practice, taking limits of vector-valued functions is no more difficult than taking
limits of real-valued functions.

Definition 68: Limits of Vector-Valued Functions


Let I be an open interval containing c , and let ⇀
r (t) be a vector-valued function defined on I , except possibly at c . The limit of


r (t) , as t approaches c , is L , expressed as


lim r (t) = L, (11.2.1)
t→c


means that given any ϵ > 0 , there exists a δ > 0 such that for all t ≠ c , if |t − c| < δ , we have ∥ r (t) − L∥ < ϵ. ⇀

Note how the measurement of distance between real numbers is the absolute value of their difference; the measure of distance
between vectors is the vector norm, or magnitude, of their difference.

theorem 89: Limits of Vector-Valued Functions


1. Let ⇀
r (t) = ⟨ f (t), g(t) ⟩ be a vector-valued function in R defined on an open interval I containing c . Then
2


lim r (t) = ⟨lim f (t) , lim g(t)⟩. (11.2.2)
t→c t→c t→c

2. Let ⇀
r (t) = ⟨ f (t), g(t), h(t) ⟩ be a vector-valued function in R defined on an open interval I containing c . Then
3


lim r (t) = ⟨lim f (t) , lim g(t) , lim h(t)⟩ (11.2.3)
t→c t→c t→c t→c

Theorem 89 states that we compute limits component-wise.

Example 11.2.1: Finding limits of vector-valued functions

Let ⇀
r (t) = ⟨
sin t

t
2
, t − 3t + 3, cos t⟩. Find lim t→0

r (t) .

Solution
We apply the theorem and compute limits component-wise.
sin t
⇀ 2
lim r (t) = ⟨lim , lim t − 3t + 3 , lim cos t⟩
t→0 t→0 t t→0 t→0

= ⟨1, 3, 1⟩.

Continuity
Definition 69: Continuity of Vector-Valued Functions
Let ⇀
r (t) be a vector-valued function defined on an open interval I containing c .
1. r (t) is continuous at c if lim

r (t) = r(c) .
t→c

2. If r (t) is continuous at all c in I , then r (t) is continuous on I .


⇀ ⇀

We again have a theorem that lets us evaluate continuity component-wise.

11.2.1 https://math.libretexts.org/@go/page/4221
THEOREM 90: Continuity of Vector-Valued Functions
Let r (t) be a vector-valued function defined on an open interval I containing c .
⇀ ⇀
r (t) is continuous at c if, and only if, each of
its component functions is continuous at c .

Example 11.2.2: Evaluating continuity of vector-valued functions

Let ⇀
r (t) = ⟨
sin t

t
2
, t − 3t + 3, cos t⟩. Determine whether ⇀
r is continuous at t = 0 and t = 1 .
Solution
While the second and third components of r (t) are defined at t = 0 , the first component, (sin t)/t , is not. Since the first

component is not even defined at t = 0 , r (t) is not defined at t = 0 , and hence it is not continuous at t = 0 .

At t = 1 each of the component functions is continuous. Therefore ⇀


r (t) is continuous at t = 1 .

Derivatives
Consider a vector-valued function r defined on an open interval I containing t and t . We can compute the displacement of r on

0 1

[ t , t ], as shown in Figure 11.2.1a. Recall that dividing the displacement vector by t − t


0 1 gives the average rate of change on
1 0

[ t , t ], as shown in 11.2.1b.
0 1

Figure 11.2.1 : Illustrating displacement, leading to an understanding of the derivative of vector-valued functions.
The derivative of a vector-valued function is a measure of the instantaneous rate of change, measured by taking the limit as the
length of [t , t ] goes to 0. Instead of thinking of an interval as [t , t ], we think of it as [c, c + h] for some value of h (hence the
0 1 0 1

interval has length h ). The average rate of change is


⇀ ⇀
r (c + h) − r (c)
(11.2.4)
h

for any value of h ≠ 0 . We take the limit as h → 0 to measure the instantaneous rate of change; this is the derivative of ⇀
r .

Definition 70: Derivative of a Vector-Valued Function


Let ⇀
r (t) be continuous on an open interval I containing c .
1. The derivative of ⇀
r at t = c is
⇀ ⇀
r (c + h) − r (c)
⇀′
r (c) = lim . (11.2.5)
h→0 h

2. The derivative of ⇀
r is
⇀ ⇀
r (t + h) − r (t)
⇀′
r (t) = lim . (11.2.6)
h→0 h

Note: Alternate notations for the derivative of ⇀


r include:

⇀′
d ⇀
dr
r (t) = ( r (t) ) = . (11.2.7)
dt dt

If a vector-valued function has a derivative for all c in an open interval I , we say that ⇀
r (t) is differentiable on I .
Once again we might view this definition as intimidating, but recall that we can evaluate limits component-wise. The following
theorem verifies that this means we can compute derivatives component-wise as well, making the task not too difficult.

11.2.2 https://math.libretexts.org/@go/page/4221
theorem 91: Derivatives of Vector-Valued Functions
1. Let ⇀
r (t) = ⟨ f (t), g(t) ⟩ . Then
⇀′ ′ ′
r (t) = ⟨ f (t), g (t) ⟩. (11.2.8)

2. Let ⇀
r (t) = ⟨ f (t), g(t), h(t) ⟩ . Then
⇀′ ′ ′ ′
r (t) = ⟨ f (t), g (t), h (t) ⟩. (11.2.9)

Example 11.2.3: Derivatives of vector-valued functions

Let ⇀ 2
r (t) = ⟨t , t⟩ .
⇀′
1. Sketch r (t) and r (t) on the same axes.

2. Compute r (1) and sketch this vector with its initial point at the origin and at
⇀′ ⇀
r (1) .
Solution
1. Theorem 91 allows us to compute derivatives component-wise, so
⇀′
r (t) = ⟨2t, 1⟩. (11.2.10)

⇀′
r (t) and r (t) are graphed together in Figure 11.9(a). Note how plotting the two of these together, in this way, is not very

illuminating. When dealing with real-valued functions, plotting f (x) with f (x) gave us useful information as we were

able to compare f and f at the same x-values. When dealing with vector-valued functions, it is hard to tell which points on

⇀′
the graph of r correspond to which points on the graph of r . ⇀

⇀′
2. We easily compute r (1) = ⟨2, 1⟩, which is drawn in Figure 11.2.2a with its initial point at the origin, as well as at
r (1) = ⟨1, 1⟩. These are sketched in Figure 11.2.2b.

Figure 11.2.2 : Graphing the derivative of a vector-valued function in Example 11.2.3

Example 11.2.4: Derivatives of vector-valued functions


⇀′ ⇀′ ⇀′
Let ⇀
r (t) = ⟨cos t, sin t, t⟩ . Compute r (t) and r (π/2) . Sketch r (π/2) with its initial point at the origin and at ⇀
r (π/2) .
Solution
⇀′ ⇀′ ⇀′
We compute r as . At t = π/2 , we have r (π/2) = ⟨−1, 0, 1⟩. Figure
r (t) = ⟨− sin t, cos t, 1⟩ 11.2.3 shows a graph of
⇀′
r (t) , with r (π/2) plotted with its initial point at the origin and at r (π/2).
⇀ ⇀

11.2.3 https://math.libretexts.org/@go/page/4221
Figure 11.2.3 : Viewing a vector-valued function and its derivative at one point.

In Examples 11.2.3 and 11.2.4, sketching a particular derivative with its initial point at the origin did not seem to reveal anything
significant. However, when we sketched the vector with its initial point on the corresponding point on the graph, we did see
something significant: the vector appeared to be tangent to the graph. We have not yet defined what "tangent'' means in terms of
curves in space; in fact, we use the derivative to define this term.

Definition 71: Tangent Vector, Tangent Line



⇀′
Let ⇀
r (t) be a differentiable vector-valued function on an open interval I containing c , where r (c) ≠ 0 .
⇀′
1. A vector v is tangent to the graph of r (t) at t = c if v is parallel to r (c).
⇀ ⇀ ⇀

⇀′
2. The tangent line} to the graph of r (t) at t = c is the line through r (c) with direction parallel to
⇀ ⇀
r (c) . An equation of the
tangent line is

⇀ ⇀′
ℓ (t) = r (c) + t r (c). (11.2.11)

Example 11.2.5: Finding tangent lines to curves in space

Let ⇀ 2
r (t) = ⟨t, t , t ⟩
3
on [−1.5, 1.5]. Find the vector equation of the line tangent to the graph of ⇀
r at t = −1 .
Solution
To find the equation of a line, we need a point on the line and the line's direction. The point is given by ⇀
r (−1) = ⟨−1, 1, −1⟩ .
(To be clear, ⟨−1, 1, −1⟩ is a vector, not a point, but we use the point "pointed to'' by this vector.)
⇀′ ⇀′ ⇀′
The direction comes from r (−1) . We compute, component-wise, 2
r (t) = ⟨1, 2t, 3 t ⟩ . Thus r (−1) = ⟨1, −2, 3⟩ .

The vector equation of the line is ℓ(t) = ⟨−1, 1, −1⟩ + t⟨1, −2, 3⟩ . This line and ⇀
r (t) are sketched, from two perspectives, in
Figure 11.2.4a and 11.2.4b.

Figure 11.2.4 : Graphing a curve in space with its tangent line.

Example 11.2.6: Finding tangent lines to curves

Find the equations of the lines tangent to ⇀ 3 2


r (t) = ⟨t , t ⟩ at t = −1 and t = 0 .

Solution
⇀′
We find that 2
r (t) = ⟨3 t , 2t⟩ . At t = −1 , we have

11.2.4 https://math.libretexts.org/@go/page/4221
⇀ ⇀′
r (−1) = ⟨−1, 1⟩ and r (−1) = ⟨3, −2⟩,

so the equation of the line tangent to the graph of ⇀


r (t) at t = −1 is

ℓ(t) = ⟨−1, 1⟩ + t⟨3, −2⟩.

This line is graphed with ⇀


r (t) in Figure 11.2.5.

Figure 11.2.5 : Graphing ⇀


r (t) and its tangent line in Example 11.2.6 .

⇀′
At t = 0 , we have r (0) = ⟨0, 0⟩ = 0 ! This implies that the tangent line "has no direction.'' We cannot apply Definition 71,
hence cannot find the equation of the tangent line.


⇀′
We were unable to compute the equation of the tangent line to r (t) = ⟨t , t ⟩ at t = 0 because r (0) = 0 . The graph in Figure
⇀ 3 2

11.12 shows that there is a cusp at this point. This leads us to another definition of smooth, previously defined by Definition 46 in
Section 9.2.

Definition 72: Smooth Vector-Valued Functions



⇀′
Let ⇀
r (t) be a differentiable vector-valued function on an open interval I . ⇀
r (t) is smooth on I if r (t) ≠ 0 on I .

Having established derivatives of vector-valued functions, we now explore the relationships between the derivative and other
vector operations. The following theorem states how the derivative interacts with vector addition and the various vector products.

THEOREM 92: Properties of Derivatives of Vector-Valued Functions


Let ⇀
r and ⇀
s be differentiable vector-valued functions, let f be a differentiable real-valued function, and let c be a real number.
⇀′
1. d

dt
⇀ ⇀
( r (t) ± s (t)) = r (t) ± s
⇀ ′
(t)

2. d

dt

(c r (t)) = c r (t)
⇀′

⇀′
3. d

dt
⇀ ′ ⇀
(f (t) r (t)) = f (t) r (t) + f (t) r (t) Product Rule
⇀′
4. d

dt
⇀ ⇀ ⇀
( r (t) ⋅ s (t)) = r (t) ⋅ s (t) + r (t) ⋅ s
⇀ ⇀ ′
(t) Product Rule
⇀′
5. d

dt
⇀ ⇀
( r (t) × s (t)) = r (t) × s (t) + r (t) × s
⇀ ⇀ ⇀ ′
(t) Product Rule
⇀′
6. d

dt

( r (f (t))) = r (f (t))f (t)

Chain Rule

Example 11.2.7: Using derivative properties of vector-valued functions

Let ⇀
r (t) = ⟨t, t
2
− 1⟩ and let u (t) be the unit vector that points in the direction of
⇀ ⇀
r (t) .
1. Graph r (t) and u (t) on the same axes, on [−2, 2].
⇀ ⇀

2. Find u (t) and sketch u (−2), u (−1) and u (0). Sketch each with initial point the corresponding point on the graph of
⇀ ′ ⇀ ′ ⇀ ′ ⇀ ′

u.

Solution
1. To form the unit vector that points in the direction of ⇀
r , we need to divide ⇀
r (t) by its magnitude.

11.2.5 https://math.libretexts.org/@go/page/4221
−−−−−−−−−−−
⇀ 2 2 2 ⇀
1 2
∥ r (t)∥ = √ t + (t − 1) ⇒ u (t) = − − −−−− −−−−− ⟨t, t − 1⟩. (11.2.12)
2 2 2
√ t + (t − 1 )


r (t) and u (t) are graphed in Figure 11.13. Note how the graph of u (t) forms part of a circle; this must be the case, as the
⇀ ⇀

length of u (t) is 1 for all t .


2. To compute u (t) , we use Theorem 92, writing


⇀ ′

⇀ ⇀
1 2 2 2 −1/2
u (t) = f (t) r (t), where f (t) = − − −−−− −−−−− = (t + (t − 1 ) ) . (11.2.13)
2 2 2
√ t + (t − 1 )

(We could write


2

t t −1
u (t) = ⟨ − − −−−−−−−− −, − − −−−−−−− − −⟩ (11.2.14)
2 2 2
√ t + (t − 1 ) √ t + (t2 − 1 )2
2

and then take the derivative. It is a matter of preference; this latter method requires two applications of the Quotient Rule
where our method uses the Product and Chain Rules.)

We find f ′
(t) using the Chain Rule:


1 2 2 2 −3/2 2
f (t) =− (t + (t − 1) ) (2t + 2(t − 1)(2t))
2
2
2t(2 t − 1)
=−
− − −−−−−−−− − 3
2 2 2
2(√ t + (t − 1 ) )

We now find u ⇀ ′
(t) using part 3 of Theorem 92:
⇀ ′ ′ ⇀ ⇀ ′
u (t) = f (t) u (t) + f (t) u (t)

2
2t(2 t − 1) 1
2
=− ⟨t, t − 1⟩ + − −−− −− −−− − − ⟨1, 2t⟩.
− − −−−−−−−− − 3
2 2
2(√ t + (t − 1 )
2
) √ t2 + (t2 − 1 )2

This is admittedly very "messy;'' such is usually the case when we deal with unit vectors. We can use this formula to
compute u (−2), u (−1) and u (0):
⇀ ⇀ ⇀


15 10
u (−2) = ⟨− −−, − −− ⟩ ≈ ⟨−0.320, −0.213⟩
13 √13 13 √13

u (−1) = ⟨0, −2⟩

u (0) = ⟨1, 0⟩

Figure 11.2.6 : Graphing in Example 11.2.7


⇀ ⇀
r (t) and u (t)

11.2.6 https://math.libretexts.org/@go/page/4221
Each of these is sketched in Figure 11.2.6. Note how the length of the vector gives an indication of how quickly the circle is
being traced at that point. When t = −2 , the circle is being drawn relatively slow; when t = −1 , the circle is being traced
much more quickly.

It is a basic geometric fact that a line tangent to a circle at a point P is perpendicular to the line passing through the center of the
circle and P . This is illustrated in Figure 11.14; each tangent vector is perpendicular to the line that passes through its initial point
and the center of the circle. Since the center of the circle is the origin, we can state this another way: u (t) is orthogonal to u (t) . ⇀ ′ ⇀

Recall that the dot product serves as a test for orthogonality: if ⇀


u⋅ v =0

, then ⇀
u is orthogonal to ⇀
v . Thus in the above example,
u (t) ⋅ u (t) = 0 .
⇀ ⇀ ′

This is true of any vector-valued function that has a constant length, that is, that traces out part of a circle. It has important
implications later on, so we state it as a theorem (and leave its formal proof as an Exercise.)

THEOREM 93 Vector-Valued Functions of Constant Length


Let r (t) be a differentiable vector-valued function on an open interval I of constant length. That is, ∥ r (t)∥ = c for all t in
⇀ ⇀
I
⇀′
(equivalently, r (t) ⋅ r (t) = c for all t in I ). Then r (t) ⋅ r (t) = 0 for all t in I .
⇀ ⇀ 2 ⇀

Integration
Indefinite and definite integrals of vector-valued functions are also evaluated component-wise.

THEOREM 94: Indefinite and Definite Integrals of Vector-Valued Functions


Let ⇀
r (t) = ⟨f (t), g(t)⟩ be a vector-valued function in R . 2

1. ∫ ⇀
r (t) dt = ⟨∫ f (t) dt, ∫ g(t) dt⟩
b ⇀ b b
2. ∫ a
r (t) dt = ⟨∫
a
f (t) dt, ∫
a
g(t) dt⟩

A similar statement holds for vector-valued functions in R . 3

Example 11.2.8: Evaluating a definite integral of a vector-valued function


1 ⇀
Let ⇀
r (t) = ⟨e
2t
, sin t⟩ . Evaluate ∫ 0
r (t) dt .
Solution
We follow Theorem 94.
1 1
⇀ 2t
∫ r (t) dt = ∫ ⟨e , sin t⟩ dt
0 0

1 1
2t
= ⟨∫ e dt , ∫ sin t dt⟩
0 0

1 1 1
2t ∣ ∣
=⟨ e , − cos t ⟩
∣0 ∣0
2
1 2
=⟨ (e − 1) , − cos(1) + 1⟩
2

≈ ⟨3.19, 0.460⟩.

Example 11.2.9: Solving an initial value problem


⇀′′
Let r (t) = ⟨2, cos t, 12t⟩ . Find ⇀
r (t) where:
1. ⇀
r (0) = ⟨−7, −1, 2⟩ and
⇀′
2. r (0) = ⟨5, 3, 0⟩.

Solution
⇀′′ ⇀′
Knowing r (t) = ⟨2, cos t, 12t⟩ , we find r (t) by evaluating the indefinite integral.

11.2.7 https://math.libretexts.org/@go/page/4221
⇀′′
∫ r (t)dt = ⟨∫ 2dt , ∫ cos tdt , ∫ 12tdt⟩

2
= ⟨2t + C1 , sin t + C2 , 6 t + C3 ⟩

2
= ⟨2t, sin t, 6 t ⟩ + ⟨C1 , C2 , C3 ⟩

2
= ⟨2t, sin t, 6 t ⟩ + C1 .


Note how each indefinite integral creates its own constant which we collect as one constant vector C1 . Knowing

⇀′
r (0) = ⟨5, 3, 0⟩ allows us to solve for C : 1


⇀′ 2
r (t) = ⟨2t, sin t, 6 t ⟩ + C1

⇀′
r (0) = ⟨0, 0, 0⟩ + C1

⟨5, 3, 0⟩ = C1 .

⇀′
So 2
r (t) = ⟨2t, sin t, 6 t ⟩ + ⟨5, 3, 0⟩ = ⟨2t + 5, sin t + 3, 6 t ⟩
2
. To find ⇀
r (t) , we integrate once more.

⇀′ 2
∫ r (t) dt = ⟨∫ 2t + 5 dt, ∫ sin t + 3 dt, ∫ 6 t dt⟩


2 3
= ⟨t + 5t, − cos t + 3t, 2 t ⟩ + C2 .


With ⇀
r (0) = ⟨−7, −1, 2⟩ , we solve for C : 2


⇀ 2 3
r (t) = ⟨t + 5t, − cos t + 3t, 2 t ⟩ + C2


r (0) = ⟨0, −1, 0⟩ + C2

⟨−7, −1, 2⟩ = ⟨0, −1, 0⟩ + C2

⟨−7, 0, 2⟩ = C2 .

So ⇀
r (t) = ⟨t
2 3
+ 5t, − cos t + 3t, 2 t ⟩ + ⟨−7, 0, 2⟩ = ⟨t
2
+ 5t − 7, − cos t + 3t, 2 t
3
+ 2⟩.

What does the integration of a vector-valued function mean? There are many applications, but none as direct as "the area under the
curve'' that we used in understanding the integral of a real-valued function.
A key understanding for us comes from considering the integral of a derivative:
b
b
⇀′ ⇀ ∣ ⇀ ⇀
∫ r (t)dt = r (t) = r (b) − r (a). (11.2.15)

a
a

Integrating a rate of change function gives displacement.


Noting that vector-valued functions are closely related to parametric equations, we can describe the arc length of the graph of a
vector-valued function as an integral. Given parametric equations x = f (t) , y = g(t) , the arc length on [a, b] of the graph is
b −−−−−−−−−−−
′ 2 ′ 2
Arc Length = ∫ √ f (t) + g (t) dt, (11.2.16)
a

−−−−−−−−−−− ⇀′
as stated in Theorem 82 in Section 9.3. If r (t) = ⟨f (t), g(t)⟩ , note that √f (t) + g (t) = ∥ r (t)∥ . Therefore we can express
⇀ ′ 2 ′ 2

the arc length of the graph of a vector-valued function as an integral of the magnitude of its derivative.

THEOREM 95: Arc Length of a Vector-Valued Function


⇀′
Let ⇀
r (t) be a vector-valued function where r (t) is continuous on [a, b]. The arc length L of the graph of ⇀
r (t) is
b
⇀′
L =∫ ∥ r (t)∥ dt. (11.2.17)
a

Note that we are actually integrating a scalar-function here, not a vector-valued function.

11.2.8 https://math.libretexts.org/@go/page/4221
The next section takes what we have established thus far and applies it to objects in motion. We will let ⇀
r (t) describe the path of
⇀′ ⇀′′
an object in the plane or in space and will discover the information provided by r (t) and r (t).

This page titled 11.2: Calculus and Vector-Valued Functions is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated
by Gregory Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is
available upon request.

11.2.9 https://math.libretexts.org/@go/page/4221
11.3: The Calculus of Motion
A common use of vector--valued functions is to describe the motion of an object in the plane or in space. A position function r (t) ⇀

gives the position of an object at time t . This section explores how derivatives and integrals are used to study the motion described
by such a function.

Definition 73: Velocity, Speed and Acceleration


Let ⇀
r (t) be a position function in R or R .
2 3

Velocity, denoted ⇀
v (t) , is the instantaneous rate of position change; that is,
⇀ ⇀′
v (t) = r (t). (11.3.1)

Speed is the magnitude of velocity,



speed = ∥ v (t)∥. (11.3.2)

Acceleration, denoted ⇀
a (t) , is the instantaneous rate of velocity change; that is,
⇀ ⇀ ′ ⇀′′
a (t) = v (t) = r (t). (11.3.3)

Example 11.3.1: Finding velocity and acceleration

An object is moving with position function ⇀


r (t) = ⟨t
2
− t, t
2
+ t⟩ , −3 ≤ t ≤ 3 , where distances are measured in feet and
time is measured in seconds.
1. Find v (t) and a (t) .
⇀ ⇀

2. Sketch r (t) ; plot v (−1), a (−1), v (1) and


⇀ ⇀ ⇀ ⇀ ⇀
a (1) , each with their initial point at their corresponding point on the graph of
r (t) .

3. When is the object's speed minimized?


Solution
1. Taking derivatives, we find
⇀ ⇀′
v (t) = r (t) = ⟨2t − 1, 2t + 1⟩

and
⇀ ⇀′′
a (t) = r (t) = ⟨2, 2⟩.

Note that acceleration is constant.


2. v (−1) = ⟨−3, −1⟩ , a (−1) = ⟨2, 2⟩ ; v (1) = ⟨1, 3⟩, a (1) = ⟨2, 2⟩. These are plotted with r (t) in Figure 11.3.1a.
⇀ ⇀ ⇀ ⇀ ⇀

We can think of acceleration as "pulling'' the velocity vector in a certain direction. At t = −1 , the velocity vector points
down and to the left; at t = 1 , the velocity vector has been pulled in the ⟨2, 2⟩ direction and is now pointing up and to the
right. In Figure 11.15(b) we plot more velocity/acceleration vectors, making more clear the effect acceleration has on
velocity.
Since a (t) is constant in this example, as t grows large v (t) becomes almost parallel to a (t) . For instance, when t = 10 ,
⇀ ⇀ ⇀

v (10) = ⟨19, 21⟩, which is nearly parallel to ⟨2, 2⟩.


3. The object's speed is given by


−−−−−−−−−−−−−−−− − −− −−−
⇀ 2 2 2
∥ v (t)∥ = √ (2t − 1 ) + (2t + 1 ) = √ 8t + 2 . (11.3.4)

To find the minimal speed, we could apply calculus techniques (such as set the derivative equal to 0 and solve for t , etc.)
but we can find it by inspection. Inside the square root we have a quadratic which is minimized when t = 0 . Thus the speed

is minimized at t = 0 , with a speed of √2 ft/s.

The graph in Figure 11.3.1b also implies speed is minimized here. The filled dots on the graph are located at integer values
of t between −3 and 3. Dots that are far apart imply the object traveled a far distance in 1 second, indicating high speed;

11.3.1 https://math.libretexts.org/@go/page/4222
dots that are close together imply the object did not travel far in 1 second, indicating a low speed. The dots are closest
together near t = 0 , implying the speed is minimized near that value.

Figure 11.3.1 : Graphing the position, velocity and acceleration of an object in Example 11.3.1

Example 11.3.2: Analyzing Motion

Two objects follow an identical path at different rates on [−1, 1]. The position function for Object 1 is r (t) = ⟨t, t ⟩ ; the ⇀
1
2

position function for Object 2 is r (t) = ⟨t , t ⟩ , where distances are measured in feet and time is measured in seconds.

2
3 6

Compare the velocity, speed and acceleration of the two objects on the path.

Solution
We begin by computing the velocity and acceleration function for each object:
⇀ ⇀ 2 5
v 1 (t) = ⟨1, 2t⟩ v 2 (t) = ⟨3 t , 6 t ⟩

⇀ ⇀ 4
a 1 (t) = ⟨0, 2⟩ a 2 (t) = ⟨6t, 30 t ⟩

We immediately see that Object 1 has constant acceleration, whereas Object 2 does not.
At t = −1 , we have v (−1) = ⟨1, −2⟩ and v (−1) = ⟨3, −6⟩ ; the velocity of Object 2 is three times that of Object 1 and so

1

2
– –
it follows that the speed of Object 2 is three times that of Object 1 (3√5 ft/s compared to √5 ft/s.)

Figure 11.3.2 : Plotting velocity and acceleration vectors for Object 1 in Example 11.3.2

At t = 0 , the velocity of Object 1 is ⇀
v (1) = ⟨1, 0⟩ and the velocity of Object 2 is 0 ! This tells us that Object 2 comes to a
complete stop at t = 0 .
In Figure 11.3.2, we see the velocity and acceleration vectors for Object 1 plotted for t = −1, −1/2, 0, 1/2 and t = 1 . Note
again how the constant acceleration vector seems to "pull'' the velocity vector from pointing down, right to up, right. We could
plot the analogous picture for Object 2, but the velocity and acceleration vectors are rather large ( a (−1) = ⟨−6, 30⟩!)

2

11.3.2 https://math.libretexts.org/@go/page/4222
Instead, we simply plot the locations of Object 1 and 2 on intervals of 1/10 of a second, shown in Figure 11.3.3a and
th

11.3.3b . Note how the x-values of Object 1 increase at a steady rate. This is because the x-component of a (t) is 0; there is no

acceleration in the x-component. The dots are not evenly spaced; the object is moving faster near t = −1 and t = 1 than near
t = 0.

Figure 11.3.3 : Comparing the positions of Objects 1 and 2 in Example 11.3.2


In Figure 11.3.3b, we see the points plotted for Object 2. Note the large change in position from t = −1 to t = −0.9 ; the
object starts moving very quickly. However, it slows considerably at it approaches the origin, and comes to a complete stop at
t = 0 . While it looks like there are 3 points near the origin, there are in reality 5 points there.

Since the objects begin and end at the same location, the have the same displacement. Since they begin and end at the same
time, with the same displacement, they have they have the same average rate of change (i.e, they have the same average
velocity). Since they follow the same path, they have the same distance traveled. Even though these three measurements are the
same, the objects obviously travel the path in very different ways.

Example 11.3.3: Analyzing the motion of a whirling ball on a string

A young boy whirls a ball, attached to a string, above his head in a counter-clockwise circle. The ball follows a circular path
and makes 2 revolutions per second. The string has length 2ft.
1. Find the position function r (t) that describes this situation.

2. Find the acceleration of the ball and derive a physical interpretation of it.
3. A tree stands 10 ft in front of the boy. At what t -values should the boy release the string so that the ball hits the tree?
Solution
1. The ball whirls in a circle. Since the string is 2 ft long, the radius of the circle is 2. The position function
r (t) = ⟨2 cos t, 2 sin t⟩ describes a circle with radius 2, centered at the origin, but makes a full revolution every 2π

seconds, not two revolutions per second. We modify the period of the trigonometric functions to be 1/2 by multiplying t by
4π. The final position function is thus


r (t) = ⟨2 cos(4πt), 2 sin(4πt)⟩. (11.3.5)

(Plot this for 0 ≤ t ≤ 1/2 to verify that one revolution is made in 1/2 a second.)
2. To find a (t) , we derive r (t) twice.
⇀ ⇀

11.3.3 https://math.libretexts.org/@go/page/4222
⇀ ⇀′
v (t) = r (t) = ⟨−8π sin(4πt), 8π cos(4πt)⟩

⇀ ⇀′′ 2 2
a (t) = r (t) = ⟨−32 π cos(4πt), −32 π sin(4πt)⟩

2
= −32 π ⟨cos(4πt), sin(4πt)⟩.

Note how ⇀
a (t) is parallel to ⇀
r (t) , but has a different magnitude and points in the opposite direction. Why is this?

Recall the classic physics equation, "Force = mass × acceleration.'' A force acting on a mass induces acceleration (i.e., the
mass moves); acceleration acting on a mass induces a force (gravity gives our mass a weight). Thus force and acceleration
are closely related. A moving ball "wants'' to travel in a straight line. Why does the ball in our example move in a circle? It
is attached to the boy's hand by a string. The string applies a force to the ball, affecting it's motion: the string accelerates
the ball. This is not acceleration in the sense of "it travels faster;'' rather, this acceleration is changing the velocity of the
ball. In what direction is this force/acceleration being applied? In the direction of the string, towards the boy's hand.

The magnitude of the acceleration is related to the speed at which the ball is traveling. A ball whirling quickly is rapidly
changing direction/velocity. When velocity is changing rapidly, the acceleration must be "large.''

3. When the boy releases the string, the string no longer applies a force to the ball, meaning acceleration is 0 and the ball can
now move in a straight line in the direction of v (t) . ⇀

Let t = t be the time when the boy lets go of the string. The ball will be at r (t ), traveling in the direction of
0

0

. We
v (t0 )

want to find t so that this line contains the point (0, 10) (since the tree is 10ft directly in front of the boy).
0

There are many ways to find this time value. We choose one that is relatively simple computationally. As shown in Figure
11.3.4, the vector from the release point to the tree is ⟨0, 10⟩ − r (t ) . This line segment is tangent to the circle, which

0

means it is also perpendicular to r (t ) itself, so their dot product is 0.



0

⇀ ⇀
r (t0 ) ⋅ (⟨0, 10⟩ − r (t0 )) = 0

⟨2 cos(4π t0 ), 2 sin(4π t0 )⟩ ⋅ ⟨−2 cos(4π t0 ), 10 − 2 sin(4π t0 )⟩ = 0

2 2
−4 cos (4π t0 ) + 20 sin(4π t0 ) − 4 sin (4π t0 ) = 0

20 sin(4π t0 ) − 4 = 0

sin(4π t0 ) = 1/5

−1
4πt0 = sin (1/5)

4πt0 ≈ 0.2 + 2πn,

where n is an integer. Solving for t we have: 0

t0 ≈ 0.016 + n/2

This is a wonderful formula. Every 1/2 second after t = 0.016s the boy can release the string (since the ball makes 2
revolutions per second, he has two chances each second to release the ball).

11.3.4 https://math.libretexts.org/@go/page/4222
Figure 11.3.4 : Modeling the flight of a ball in Example 11.3.3

Example 11.3.4: Analyzing motion in space

An object moves in a spiral with position function r (t) = ⟨cos t, sin t, t⟩ , where distances are measured in meters and time is

in minutes. Describe the object's speed and acceleration at time t .

Solution
With r (t) = ⟨cos t, sin t, t⟩ , we have:


v (t) = ⟨− sin t, cos t, 1⟩ and


a (t) = ⟨− cos t, − sin t, 0⟩.

−−−−−−−−−−−−−−−−− –
The speed of the object is ∥ v (t)∥ = √(− sin t) + cos t + 1 = √2 m/min; it moves at a constant speed. Note that the object
⇀ 2 2

does not accelerate in the z -direction, but rather moves up at a constant rate of 1m/min.

The objects in Examples 11.3.3 and 11.3.4 traveled at a constant speed. That is, ∥ v (t)∥ = c for some constant c . Recall Theorem

93, which states that if a vector--valued function r (t) \ has constant length, then r (t) \ is perpendicular to its derivative:
⇀ ⇀

r (t) ⋅ r (t) = 0 . In these examples, the velocity function has constant length, therefore we can conclude that the velocity is
⇀ ⇀′

perpendicular to the acceleration: v (t)cdot a (t) = 0 . A quick check verifies this.


⇀ ⇀

There is an intuitive understanding of this. If acceleration is parallel to velocity, then it is only affecting the object's speed; it does
not change the direction of travel. (For example, consider a dropped stone. Acceleration and velocity are parallel -- straight down --
and the direction of velocity never changes, though speed does increase.) If acceleration is not perpendicular to velocity, then there
is some acceleration in the direction of travel, influencing the speed. If speed is constant, then acceleration must be orthogonal to
velocity, as it then only affects direction, and not speed.

key idea: 52 Objects With Constant Speed


If an object moves with constant speed, then its velocity and acceleration vectors are orthogonal. That is, ⇀ ⇀
v (t)cdot a (t) = 0 .

Projectile Motion
An important application of vector--valued position functions is projectile motion: the motion of objects under only the influence of
gravity. We will measure time in seconds, and distances will either be in meters or feet. We will show that we can completely
describe the path of such an object knowing its initial position and initial velocity (i.e., where it is and where it is going.)

Suppose an object has initial position r (0) = ⟨x , y ⟩ and initial velocity v (0) = ⟨v , v ⟩ . It is customary to rewrite v (0) in

0 0

x y

terms of its speed v and direction u , where u is a unit vector. Recall all unit vectors in R can be written as ⟨cos θ, sin θ⟩ , where
0
⇀ ⇀ 2

θ is an angle measure counter--clockwise from the x-axis. (We refer to θ as the angle of elevation.) Thus v (0) = v ⟨cos θ, sin θ⟩.

0

11.3.5 https://math.libretexts.org/@go/page/4222
Since the acceleration of the object is known, namely a (t) = ⟨0, −g⟩ , where

g is the gravitational constant, we can find ⇀
r (t)

knowing our two initial conditions. We first find \(\vecs v (t)):

Note

In this text we use g = 32 ft/s when using Imperial units, and g = 9.8 m/s when using SI units.

⇀ ⇀
v (t) = ∫ a (t)dt


v (t) = ∫ ⟨0, −g⟩dt



v (t) = ⟨0, −gt⟩ + C.


Knowing ⇀
v (0) = v0 ⟨cos θ, sin θ⟩ , we have C = v 0 ⟨cos t, sin t⟩ and so

v (t) = ⟨v0 cos θ, −gt + v0 sin θ⟩. (11.3.6)

We integrate once more to find ⇀


r (t) :

⇀ ⇀
r (t) = ∫ v (t)dt


r (t) = ∫ ⟨v0 cos θ, −gt + v0 sin θ⟩dt

1 ⇀
⇀ 2
r (t) = ⟨(v0 cos θ)t, − gt + (v0 sin θ)t⟩ + C.
2



Knowing r (0) = ⟨x0 , y0 ⟩, we conclude C = ⟨x0 , y0 ⟩ and


1 2
r (t) = ⟨(v0 cos θ)t + x0 , − gt + (v0 sin θ)t + y0 ⟩.
2

key idea 53: Projectile Motion


The position function of a projectile propelled from an initial position of ⇀
r0 = ⟨x0 , y0 ⟩ , with initial speed v0 , with angle of
elevation θ and neglecting all accelerations but gravity is


1 2
r (t) = ⟨(v0 cos θ)t + x0 , − gt + (v0 sin θ)t + y0 ⟩. (11.3.7)
2

Letting ⇀
v0 = v0 ⟨cos θ, sin θ⟩ , (\vecs r (t)\) can be written as


1 2 ⇀ ⇀
r (t) = ⟨0, − gt ⟩ + v 0 t + r 0 . (11.3.8)
2

We demonstrate how to use this position function in the next two examples.

Example 11.3.5: Projectile Motion

Sydney shoots her Red Ryder bb gun across level ground from an elevation of 4ft, where the barrel of the gun makes a 5 ∘

angle with the horizontal. Find how far the bb travels before landing, assuming the bb is fired at the advertised rate of 350ft/s
and ignoring air resistance.
Solution
A direct application of Key Idea 53 gives
⇀ ∘ 2 ∘
r (t) = ⟨(350 cos 5 )t, −16 t + (350 sin 5 )t + 4⟩

2
≈ ⟨346.67t, −16 t + 30.50t + 4⟩,

where we set her initial position to be ⟨0, 4⟩.


We need to find when the bb lands, then we can find where. We accomplish this by setting the y -component equal to 0 and

11.3.6 https://math.libretexts.org/@go/page/4222
solving for t :
2
−16 t + 30.50t + 4 = 0

−−−−−−−−−−−−−−−
2
−30.50 ± √ 30.50 − 4(−16)(4)

t =
−32

t ≈ 2.03s.

(We discarded a negative solution that resulted from our quadratic equation.)
We have found that the bb lands 2.03s after firing; with t = 2.03 , we find the x -component of our position function is
346.67(2.03) = 703.74ft. The bb lands about 704 feet away.

Example 11.3.6: Projectile Motion

Alex holds his sister's bb gun at a height of 3ft and wants to shoot a target that is 6ft above the ground, 25ft away. At what
angle should he hold the gun to hit his target? (We still assume the muzzle velocity is 350ft/s.)
Solution
The position function for the path of Alex's bb is
⇀ 2
r (t) = ⟨(350 cos θ)t, −16 t + (350 sin θ)t + 3⟩.

We need to find θ so that ⇀


r (t) =\langle 25,6\rangle\) for some value of t . That is, we want to find θ and t such that
2
(350 cos θ)t = 25 and − 16 t + (350 sin θ)t + 3 = 6.

This is not trivial (though not "hard''). We start by solving each equation for cos θ and sin θ , respectively.
2
25 3 + 16t
cos θ = and sin θ = .
350t 350t

Using the Pythagorean Identity cos 2


θ + sin
2
θ =1 , we have
2 2 2
25 3 + 16t
( ) +( ) =1
350t 350t

Multiply both sides by (350t) : 2

2 2 2 2 2
25 + (3 + 16 t ) = 350 t

4 2
256 t − 122, 404 t + 634 = 0.

This is a quadratic in t . That is, we can apply the quadratic formula to find t , then solve for t itself.
2 2

−−−−−−−−−−−−−−−−−−−
2
122, 404 ± √ 122, 404 − 4(256)(634)
2
t =
512

2
t = 0.0052, 478.135

t = ±0.072, ± 21.866

Clearly the negative t values do not fit our context, so we have t = 0.072 and t = 21.866. Using cos θ = 25/(350t), we can
solve for θ :
25 25
−1 −1
θ = cos ( ) and cos ( )
350 ⋅ 0.072 350 ⋅ 21.866

∘ ∘
θ = 7.03 and 89.8 .

Alex has two choices of angle. He can hold the rifle at an angle of about 7 with the horizontal and hit his target 0.07s after

firing, or he can hold his rifle almost straight up, with an angle of 89.8 , where he'll hit his target about 22s later. The first

option is clearly the option he should choose.

11.3.7 https://math.libretexts.org/@go/page/4222
Distance Traveled
Consider a driver who sets her cruise--control to 60mph, and travels at this speed for an hour. We can ask:
1. How far did the driver travel?
2. How far from her starting position is the driver?
The first is easy to answer: she traveled 60 miles. The second is impossible to answer with the given information. We do not know
if she traveled in a straight line, on an oval racetrack, or along a slowly--winding highway.
This highlights an important fact: to compute distance traveled, we need only to know the speed, given by ∥ v (t)∥. ⇀

theorem 96: Distance Traveled


Let ⇀
v (t) be a velocity function for a moving object. The distance traveled by the object on [a, b] is:
b

distance traveled = ∫ ∥ v (t)∥dt. (11.3.9)
a

Note that this is just a restatement of Theorem 95: arc length is the same as distance traveled, just viewed in a different context.

Example 11.3.7: Distance Traveled, Displacement, and Average Speed

A particle moves in space with position function ⇀ 2


r (t) = ⟨t, t , sin(πt)⟩ on [−2, 2] , where t is measured in seconds and
distances are in meters. Find:
1. The distance traveled by the particle on [−2, 2].
2. The displacement of the particle on [−2, 2].
3. The particle's average speed.
Solution
1. We use Theorem 96 to establish the integral:
2

distance traveled = ∫ ∥ v (t)∥dt
−2

2 −−−−−−−−−−−−−−−−−−
2 2 2
=∫ √ 1 + (2t) +π cos (πt) dt.
−2

This cannot be solved in terms of elementary functions so we turn to numerical integration, finding the distance to be
12.88m.
2. The displacement is the vector
⇀ ⇀
r (2) − r (−2) = ⟨2, 4, 0⟩ − ⟨−2, 4, 0⟩ = ⟨4, 0, 0⟩.

That is, the particle ends with an x-value increased by 4 and with y - and z -values the same (see Figure 11.3.5).
3. We found above that the particle traveled 12.88m over 4 seconds. We can compute average speed by dividing: 12.88/4 =
3.22m/s.

We should also consider Definition 22 of Section 5.4, which says that the average value of a function f on [a, b] is
1 b

a
f (x) dx . In our context, the average value of the speed is
b −a

2
1 1

average speed = ∫ ∥ v (t)∥dt ≈ 12.88 = 3.22m/s.
2 − (−2) −2 4

Note how the physical context of a particle traveling gives meaning to a more abstract concept learned earlier.

11.3.8 https://math.libretexts.org/@go/page/4222
Figure 11.3.5 : The path of the particle in Example 11.3.7

In Definition 22 of Chapter 5 we defined the average value of a function f (x) on [a, b] to be


b
1
∫ f (x)dx. (11.3.10)
b −a a

Note how in Example 11.3.7 we computed the average speed as


2
distance traveled 1 ⇀
= ∫ ∥ v (t)∥dt; (11.3.11)
travel time 2 − (−2) −2

that is, we just found the average value of ∥ v (t)∥ on [−2, 2].

Likewise, given position function ⇀


r (t) , the average velocity on [a, b] is
b ⇀ ⇀
displacement 1 r (b) − r (a)
⇀ ′
= ∫ r (t)dt = ; (11.3.12)
travel time b −a a
b −a

that is, it is the average value of ⇀ ′


r (t) , or ⇀
v (t) , on [a, b].

KEY IDEA 54: Average Speed, Average Velocity


Let ⇀
r (t) be a continuous position function on an open interval I containing a < b .
The average speed is:
b ⇀ b
distance traveled ∫ ∥ v (t)∥dt 1
a ⇀
= = ∫ ∥ v (t)∥dt. (11.3.13)
travel time b −a b −a a

The average velocity is:


b ⇀ ′
b
displacement ∫ r (t)dt 1
a ⇀ ′
= = ∫ r (t)dt. (11.3.14)
travel time b −a b −a a

The next two sections investigate more properties of the graphs of vector--valued functions and we'll apply these new ideas to what
we just learned about motion.

This page titled 11.3: The Calculus of Motion is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory
Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

11.3.9 https://math.libretexts.org/@go/page/4222
11.4: Unit Tangent and Normal Vectors
Unit Tangent Vector
⇀′
Given a smooth vector-valued function r (t) , we defined in Definition 71 that any vector parallel to r (t ) is tangent to the graph

0

of r (t) at t = t . It is often useful to consider just the direction of r (t) and not its magnitude. Therefore we are interested in the

0
⇀′

⇀′
unit vector in the direction of r (t) . This leads to a definition.

Definition 11.4.1: Unit Tangent Vector



Let ⇀
r (t) be a smooth function on an open interval I . The unit tangent vector T(t) is
⇀ 1 ⇀′
T(t) = r (t). (11.4.1)
⇀′
∥ r (t)∥

Example 11.4.1: Computing the unit tangent vector


⇀ ⇀ ⇀
Let ⇀
r (t) = ⟨3 cos t, 3 sin t, 4t⟩ . Find T(t) and compute T(0) and T(1).
Solution

We apply Definition 11.4.1 to find T(t) .
⇀ 1
⇀′
T(t) = r (t)
⇀′
∥ r (t)∥

1
= −−−−−−−−−−−−−−−−−−−−− ⟨−3 sin t, 3 cos t, 4⟩
2 2 2
√ (−3 sin t) + (3 cos t) +4

3 3 4
= ⟨− sin t, cos t, ⟩.
5 5 5

⇀ ⇀
We can now easily compute T(0) and T(1):
⇀ 3 4
T(0) = ⟨0, , ⟩
5 5

⇀ 3 3 4
T(1) = ⟨− sin 1, cos 1, ⟩ ≈ ⟨−0.505, 0.324, 0.8⟩.
5 5 5

These are plotted in Figure 11.4.1 with their initial points at ⇀


r (0) and ⇀
r (1) , respectively. (They look rather "short'' since they
are only length 1.)

Figure 11.4.1 : Plotting unit tangent vectors in Example 11.4.1 .



The unit tangent vector T(t) always has a magnitude of 1, though it is sometimes easy to doubt that is true. We can help

solidify this thought in our minds by computing ∥T(1)∥:
−−−−−−−−−−−−−−−−−−−−

2 2 2
∥ T(1)∥ ≈ √ (−0.505 ) + 0.324 + 0.8 = 1.000001.

11.4.1 https://math.libretexts.org/@go/page/4223

We have rounded in our computation of T(1), so we don't get 1 exactly. We leave it to the reader to use the exact

representation of T(1) to verify it has length 1.

⇀′
In many ways, the previous example was "too nice.'' It turned out that r (t) was always of length 5. In the next example the length
⇀′
of r (t) is variable, leaving us with a formula that is not as clean.

Example 11.4.2: Computing the unit tangent vector


⇀ ⇀ ⇀
Let ⇀ 2
r (t) = ⟨t − t, t
2
+ t⟩ . Find T(t) and compute T(0) and T(1).
Solution
⇀′
We find r (t) = ⟨2t − 1, 2t + 1⟩ , and
−−−−−−−−−−−−−−−− − −− −−−
⇀′ 2 2 2
∥ r (t)∥ = √ (2t − 1 ) + (2t + 1 ) = √ 8t + 2 . (11.4.2)

Therefore
⇀ 1 2t − 1 2t + 1
T(t) = ⟨2t − 1, 2t + 1⟩ = ⟨ , ⟩. (11.4.3)
− − −−−− − − −−−− − − −−−−
√ 8 t2 + 2 √ 8 t2 + 2 √ 8 t2 + 2

⇀ – – ⇀ −− −−
When t = 0 , we have T(0) = ⟨−1/√2, 1/√2⟩; when t = 1 , we have T(1) = ⟨1/√10, 3/√10⟩. We leave it to the reader to
verify each of these is a unit vector. They are plotted in Figure 11.4.2.

Figure 11.4.2 : Plotting unit tangent vectors in Example 11.4.2 .

Unit Normal Vector


Just as knowing the direction tangent to a path is important, knowing a direction orthogonal to a path is important. When dealing
with real-valued functions, we defined the normal line at a point to the be the line through the point that was perpendicular to the
tangent line at that point. We can do a similar thing with vector-valued functions. Given r (t) in R , we have 2 directions ⇀ 2

perpendicular to the tangent vector, as shown in Figure 11.4.3. It is good to wonder "Is one of these two directions preferable over
the other?''

Figure 11.4.3 : Given a direction in the plane, there are always two directions orthogonal to it.
Given r (t) in R , there are infinite vectors orthogonal to the tangent vector at a given point. Again, we might wonder "Is one of
⇀ 3

these infinite choices preferable over the others? Is one of these the "right' choice?''
The answer in both R and R is "Yes, there is one vector that is not only preferable, it is the "right' one to choose.'' Recall
2 3


⇀′
Theorem 93, which states that if r (t) has constant length, then r (t) is orthogonal to r (t) for all t . We know T(t) , the unit
⇀ ⇀

⇀ ⇀
tangent vector, has constant length. Therefore T(t) is orthogonal to T (t) . ′

11.4.2 https://math.libretexts.org/@go/page/4223

⇀′
We'll see that T (t) is more than just a convenient choice of vector that is orthogonal to

r (t) ; rather, it is the "right'' choice. Since
all we care about is the direction, we define this newly found vector to be a unit vector.

⇀ ⇀
Note: T(t) is a unit vector, by definition. This does not imply that T ′
(t) is also a unit vector.

Definition 11.4.2: Unit Normal Vector



Let r (t) be a vector-valued function where the unit tangent vector,

T(t) , is smooth on an open interval I . The unit normal

vector N(t) is
⇀ 1 ⇀

N(t) = T (t). (11.4.4)


∥ T (t)∥

Example 11.4.3: Computing the unit normal vector


⇀ ⇀
Let ⇀
r (t) = ⟨3 cos t, 3 sin t, 4t⟩ as in Example 11.4.1. Sketch both T(π/2) and N(π/2) with initial points at ⇀
r (π/2) .
Solution

In Example 11.4.1, we found T(t) = ⟨(−3/5) sin t, (3/5) cos t, 4/5⟩. Therefore


3 3 ⇀

3
T (t) = ⟨− cos t, − sin t, 0⟩ and ∥ T (t)∥ = . (11.4.5)
5 5 5

Thus


⇀ T (t)
N(t) = = ⟨− cos t, − sin t, 0⟩. (11.4.6)
3/5

⇀ ⇀
We compute T(π/2) = ⟨−3/5, 0, 4/5⟩ and N(π/2) = ⟨0, −1, 0⟩. These are sketched in Figure 11.4.4.

Figure 11.4.4 : Plotting unit tangent and normal vectors in Example 11.4.3 .


The previous example was once again "too nice.'' In general, the expression for T(t) contains fractions of square-roots, hence the

expression of T (t) is very messy. We demonstrate this in the next example.

Example 11.4.4: Computing the unit normal vector



Let ⇀
r (t) = ⟨t
2 2
− t, t + t⟩ as in Example 11.4.2 . Find N(t) and sketch ⇀
r (t) with the unit tangent and normal vectors at
t = −1, 0 and 1.

Solution
In Example 11.4.2, we found
⇀ 2t − 1 2t + 1
T(t) = ⟨ − − −−−−, − − −−−− ⟩. (11.4.7)
√ 8 t2 + 2 √ 8 t2 + 2


Finding T ′
(t) requires two applications of the Quotient Rule:

11.4.3 https://math.libretexts.org/@go/page/4223
− − −−−− 1
2 −1/2
√ 8 t2 + 2 (2) − (2t − 1) ( (8 t + 2 ) (16t))
2

T (t) = ⟨ ,
2
8t +2

− − −−−− 1
2 −1/2
√ 8 t2 + 2 (2) − (2t + 1) ( (8 t + 2 ) (16t))
2

2
8t +2

4(2t + 1) 4(1 − 2t)


=⟨ , ⟩
3/2 3/2
2 2
(8 t + 2) (8 t + 2)

⇀ ⇀
This is not a unit vector; to find N(t), we need to divide T ′
(t) by it's magnitude.
−−−−−−−−−−−−−−−−−−−−−
2 2
⇀ 16(2t + 1) 16(1 − 2t)

∥ T (t)∥ = √ +
2 3 2 3
(8 t + 2) (8 t + 2)

−−−−−−−−−−
2
16(8 t + 2)
=√
2 3
(8 t + 2)

4
= .
2
8t +2

Finally,

⇀ 1 4(2t + 1) 4(1 − 2t)


N(t) = ⟨ , ⟩
2 3/2 3/2
4/(8 t + 2) 2 2
(8 t + 2) (8 t + 2)

2t + 1 2t − 1
=⟨ ,− ⟩.
− − −−−− − − −−−−
√ 8 t2 + 2 √ 8 t2 + 2


Using this formula for N(t), we compute the unit tangent and normal vectors for t = −1, 0 and 1 and sketch them in Figure
11.4.5.

Figure 11.4.5 : Plotting unit tangent and normal vectors in Example 11.4.4 .

⇀ ⇀
The final result for N(t) in Example 11.4.4 is suspiciously similar to T(t) . There is a clear reason for this. If u = ⟨u , u ⟩ is a ⇀
1 2

unit vector in R , then the only unit vectors orthogonal to u are ⟨−u , u ⟩ and ⟨u , −u ⟩. Given T(t) , we can quickly determine
2 ⇀
2 1 2 1

N(t) if we know which term to multiply by (−1).

Consider again Figure 11.24, where we have plotted some unit tangent and normal vectors. Note how N(t) always points "inside''
the curve, or to the concave side of the curve. This is not a coincidence; this is true in general. Knowing the direction that r (t) ⇀


"turns'' allows us to quickly find N(t).

THEOREM 11.4.1: Unit Normal Vectors in R 2

⇀ ⇀
Let ⇀
r (t) be a vector-valued function in R where T
2 ′
(t) is smooth on an open interval I . Let t be in 0 I and T(t 0) = ⟨t1 , t2 ⟩

Then N(t 0) is either
⇀ ⇀
N(t0 ) = ⟨−t2 , t1 ⟩ or N(t0 ) = ⟨t2 , −t1 ⟩, (11.4.8)

whichever is the vector that points to the concave side of the graph of ⇀
r .

11.4.4 https://math.libretexts.org/@go/page/4223
Application to Acceleration
Let r (t) be a position function. It is a fact (stated later in Theorem 11.4.2) that acceleration, \vecs a (t), lies in the plane defined by

⇀ ⇀
T and N. That is, there are scalars a and a such that
T N

⇀ ⇀

a (t) = aT T(t) + aN N(t). (11.4.9)

The scalar a measures "how much'' acceleration is in the direction of travel, that is, it measures the component of acceleration that
T

affects the speed. The scalar a measures "how much'' acceleration is perpendicular to the direction of travel, that is, it measures
N

the component of acceleration that affects the direction of travel.



We can find a using the orthogonal projection of
T

a (t) onto T(t) (review Definition 59 in Section 10.3 if needed).
⇀ ⇀ ⇀
Recalling that since T(t) is a unit vector, T(t) ⋅ T(t) = 1 , so we have


a (t) ⋅ T(t) ⇀ ⇀ ⇀
⇀ ⇀
proj T (t) a (t) = T(t) = ( a (t) ⋅ T(t)) T(t). (11.4.10)
⇀ ⇀
T(t) ⋅ T(t) 
aT

⇀ ⇀ ⇀
Thus the amount of ⇀
a (t) in the direction of T(t) is a T

= a (t) ⋅ T(t) . The same logic gives a N

= a (t) ⋅ N(t) .

While this is a fine way of computing a , there are simpler ways of finding
T aN (as finding N itself can be complicated). The
following theorem gives alternate formulas for a and a . T N

Note: Keep in mind that both a and a are functions of t ; that is, the scalar changes depending on t . It is convention to drop the "
T N

(t) '' notation from a (t) and simply write a .


T T

⇀ ⇀
THEOREM 11.4.2: Acceleration in the Plane Defined by T and N
⇀ ⇀
Let r (t) be a position function with acceleration a (t) and unit tangent and normal vectors T(t) and N(t). Then
⇀ ⇀ ⇀
a (t) lies in
⇀ ⇀
the plane defined by T(t) and N(t); that is, there exists scalars a and a such that T N

⇀ ⇀

a (t) = aT T(t) + aN N(t). (11.4.11)

Moreover,


⇀ d ⇀
aT = a (t) ⋅ T(t) = (∥ v (t)∥)
dt

⇀ ⇀

−−−−−−−−−− ∥ a (t) × v (t)∥ ⇀
⇀ ⇀ 2 2 ⇀ ′
aN = a (t) ⋅ N(t) = √ ∥ a (t)∥ −a = = ∥ v (t)∥ ∥ T (t)∥
T ⇀
∥ v (t)∥

d
Note the second formula for a : T

(∥ v (t)∥) . This measures the rate of change of speed, which again is the amount of acceleration
dt
in the direction of travel.

Example 11.4.5: Computing a and a T N

Let ⇀
r (t) = ⟨3 cos t, 3 sin t, 4t⟩ as in Examples 11.4.1 and 11.4.3. Find a and a . T N

Solution
The previous examples give ⇀
a (t) = ⟨−3 cos t, −3 sin t, 0⟩ and
⇀ 3 3 4 ⇀
T(t) = ⟨− sin t, cos t, ⟩ and N(t) = ⟨− cos t, − sin t, 0⟩. (11.4.12)
5 5 5

We can find a and a directly with dot products:


T N


⇀ 9 9
aT = a (t) ⋅ T(t) = cos t sin t − cos t sin t + 0 = 0.
5 5


⇀ 2 2
aN = a (t) ⋅ N(t) = 3 cos t + 3 sin t + 0 = 3.

11.4.5 https://math.libretexts.org/@go/page/4223
⇀ ⇀ ⇀
Thus ⇀
a (t) = 0 T(t) + 3 N(t) = 3 N(t) , which is clearly the case.
What is the practical interpretation of these numbers? aT = 0 means the object is moving at a constant speed, and hence all
acceleration comes in the form of direction change.

Example 11.4.6: Computing a and a T N

Let ⇀
r (t) = ⟨t
2
− t, t
2
+ t⟩ as in Examples 11.4.2 and 11.4.4. Find a and a . T N

Solution
The previous examples give ⇀
a (t) = ⟨2, 2⟩ and
⇀ 2t − 1 2t + 1
T(t) = ⟨ − − −−−−, − − −−−−⟩
√ 8 t2 + 2 √ 8 t2 + 2

and
⇀ 2t + 1 2t − 1
N(t) = ⟨ − −− −−−,− − − −−−− ⟩.
2
√ 8t + 2 √ 8 t2 + 2


While we can compute a using N(t), we instead demonstrate using another formula from Theorem 11.4.2.
N


⇀ 4t − 2 4t + 2 8t
aT = a (t) ⋅ T(t) = − − −−−− + − − −−−− = − − −−−−.
√ 8 t2 + 2 √ 8 t2 + 2 √ 8 t2 + 2

−−−−−−−−−−−−−−
2
−−−−−−−−−− 8t 4
⇀ 2 2
aN = √ ∥ a (t)∥ −a = √8 − ( − − −−−−) = − − −−−−
T
√ 8 t2 + 2 √ 8 t2 + 2

16 4
When t = 2 , a T = −− ≈ 2.74 and a N = −− ≈ 0.69 . We interpret this to mean that at t = 2 , the particle is accelerating
√34 √34

mostly by increasing speed, not by changing direction. As the path near t = 2 is relatively straight, this should make intuitive
sense. Figure 11.4.6 gives a graph of the path for reference.

Figure 11.4.6 : Graphing ⇀


r (t) in Example 11.4.6 .

Contrast this with t = 0 , where a T =0 and a N = 4/ √2 ≈ 2.82 . Here the particle's speed is not changing and all acceleration
is in the form of direction change.

Example 11.4.7: Analyzing projectile motion

A ball is thrown from a height of 240ft with an initial speed of 64ft/s and an angle of elevation of 30

. Find the position
function r (t) of the ball and analyze a and a .

T N

Solution
Using Key Idea 53 of Section 11.3 we form the position function of the ball:
⇀ ∘ 2 ∘
r (t) = ⟨(64 cos 30 ) t, −16 t + (64 sin 30 ) t + 240⟩, (11.4.13)

which we plot in Figure 11.4.7.

11.4.6 https://math.libretexts.org/@go/page/4223
Figure 11.4.7 : Plotting the position of a thrown ball, with 1s increments shown.

From this we find v (t) = ⟨64 cos 30
⇀ ∘
, −32t + 64 sin 30 ⟩

and ⇀
a (t) = ⟨0, −32⟩ . Computing T(t) is not difficult, and with
some simplification we find

⇀ √3 1 −t
T(t) = ⟨ − −−−− −−−−, − −−−− −−−− ⟩. (11.4.14)
√ t2 − 2t + 4 √ t2 − 2t + 4

With ⇀
a (t) as simple as it is, finding a is also simple:
T


⇀ 32t − 32
aT = a (t) ⋅ T(t) = − −−−− −−−−. (11.4.15)
√ t2 − 2t + 4

−−−−−−−−−−−

We choose to not find N(t) and find a through the formula a
N N

= √∥ a (t)∥2 − a
2
T
:
−−−−−−−−−−−−−−−−−−−
2 –
2
32t − 32 32 √3
aN = √ 32 −( − −−−− −−−−) = − −−−− −−−−. (11.4.16)
√ t2 − 2t + 4 √ t2 − 2t + 4

Figure 11.4.8 gives a table of values of a and a . When t = 0 , we see the ball's speed is decreasing; when t = 1 the speed of
T N

the ball is unchanged. This corresponds to the fact that at t = 1 the ball reaches its highest point.
After t = 1 we see that a is decreasing in value. This is because as the ball falls, it's path becomes straighter and most of the
N

acceleration is in the form of speeding up the ball, and not in changing its direction.

Figure 11.4.8 : A table of values of a and a T N in Example 11.4.7 .

Our understanding of the unit tangent and normal vectors is aiding our understanding of motion. The work in Example 11.4.7 gave
quantitative analysis of what we intuitively knew.
The next section provides two more important steps towards this analysis. We currently describe position only in terms of time. In
everyday life, though, we often describe position in terms of distance ("The gas station is about 2 miles ahead, on the left.''). The
arc length parameter allows us to reference position in terms of distance traveled.
We also intuitively know that some paths are straighter than others - and some are "curvier" than others, but we lack a measurement
of "curviness.'' The arc length parameter provides a way for us to compute curvature, a quantitative measurement of how curvy a
curve is.

This page titled 11.4: Unit Tangent and Normal Vectors is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by
Gregory Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available
upon request.

11.4.7 https://math.libretexts.org/@go/page/4223
11.5: The Arc Length Parameter and Curvature
In normal conversation we describe position in terms of both time and distance. For instance, imagine driving to visit a friend. If
she calls and asks where you are, you might answer "I am 20 minutes from your house,'' or you might say "I am 10 miles from your
house.'' Both answers provide your friend with a general idea of where you are.
Currently, our vector-valued functions have defined points with a parameter t , which we often take to represent time. Consider
Figure 11.5.1a, where r (t) = ⟨t − t, t + t⟩ is graphed and the points corresponding to t = 0, 1 and 2 are shown. Note how the
⇀ 2 2

arc length between t = 0 and t = 1 is smaller than the arc length between t = 1 and t = 2 ; if the parameter t is time and r is ⇀

position, we can say that the particle traveled faster on [1, 2] than on [0, 1].

Figure 11.5.1 : Introducing the arc length parameter.


Now consider Figure 11.5.1b, where the same graph is parametrized by a different variable s . Points corresponding to s = 0
through s = 6 are plotted. The arc length of the graph between each adjacent pair of points is 1. We can view this parameter s as
distance; that is, the arc length of the graph from s = 0 to s = 3 is 3, the arc length from s = 2 to s = 6 is 4, etc. If one wants to
find the point 2.5 units from an initial location (i.e., s = 0 ), one would compute r (2.5). This parameter s is very useful, and is

called the arc length parameter.


How do we find the arc length parameter?
Start with any parametrization of ⇀
r . We can compute the arc length of the graph of ⇀
r on the interval [0, t] with
t
⇀ ′
arc length = ∫ ∥r (u)∥du. (11.5.1)
0

We can turn this into a function: as t varies, we find the arc length s from 0 to t . This function is
t
⇀ ′
s(t) = ∫ ∥r (u)∥du. (11.5.2)
0

This establishes a relationship between s and t . Knowing this relationship explicitly, we can rewrite ⇀
r (t) as a function of s : ⇀
r (s) .
We demonstrate this in an example.

Example 11.5.1: Finding the arc length parameter

Let ⇀
r (t) = ⟨3t − 1, 4t + 2⟩ . Parametrize ⇀
r with the arc length parameter s .}
Solution
Using Equation 11.5.2, we write
t
⇀ ′
s(t) = ∫ ∥r (u)∥du.
0

11.5.1 https://math.libretexts.org/@go/page/4224
We can integrate this, explicitly finding a relationship between s and t :
t
⇀ ′
s(t) = ∫ ∥r (u)∥du
0

t
− −−−−−
2 2
=∫ √ 3 + 4 du

=∫ 5du
0

= 5t.

Since s = 5t , we can write t = s/5 and replace t in ⇀


r (t) with s/5:


3 4
r (s) = ⟨3(s/5) − 1, 4(s/5) + 2⟩ = ⟨ s − 1, s + 2⟩.
5 5

Clearly, as shown in Figure 11.5.2, the graph of r is a line, where t = 0 corresponds to the point (−1, 2). What point on the

line is 2 units away from this initial point? We find it with s(2) = ⟨1/5, 18/5⟩.

Figure 11.5.2 : Graphing vecr in Example 11.5.1 .


Is the point (1/5, 18/5)really 2 units away from (−1, 2)? We use the Distance Formula to check:
−−−−−−−−−−−−−−−−−−−−−−
2 2
1 18
d = √( − (−1)) +( − 2)
5 5

−−−−−− −
36 64
=√ +
25 25


= √4 = 2.

Yes, s(2) is indeed 2 units away, in the direction of travel, from the initial point.

Things worked out very nicely in Example 11.5.1; we were able to establish directly that s = 5t . Usually, the arc length parameter
is much more difficult to describe in terms of t , a result of integrating a square-root. There are a number of things that we can learn
about the arc length parameter from Equation 11.5.2, though, that are incredibly useful.
First, take the derivative of s with respect to t . The Fundamental Theorem of Calculus (see Theorem 39) states that
ds
′ ⇀′
= s (t) = ∥ r (t)∥. (11.5.3)
dt

Letting t represent time and r (t) represent position, we see that the rate of change of s with respect to t is speed; that is, the rate

of change of "distance traveled'' is speed, which should match our intuition.


The Chain Rule states that
⇀ ⇀
dr dr ds
= ⋅
dt ds dt

⇀′ ⇀′ ⇀′
r (t) = r (s) ⋅ ∥ r (t)∥.

⇀′
Solving for r (s) , we have

11.5.2 https://math.libretexts.org/@go/page/4224
⇀′
r (t) ⇀
⇀′
r (s) = = T(t), (11.5.4)
⇀′
∥ r (t)∥

⇀ ⇀
⇀′
where T(t) is the unit tangent vector. Equation 11.5.4 is often misinterpreted, as one is tempted to think it states r (t) = T(t) ,
but there is a big difference between r (s) and r (t) . The key to take from it is that r (s) is a unit vector. In fact, the following
⇀′ ⇀′ ⇀′

theorem states that this characterizes the arc length parameter.

theorem 99: Arc Length Parameter


⇀′
Let ⇀
r (s) be a vector-valued function. The parameter s is the arc length parameter if, and only if, ∥ r (s)∥ = 1.

Curvature
Consider points A and B on the curve graphed in Figure 11.5.3a. One can readily argue that the curve curves more sharply at A

than at B . It is useful to use a number to describe how sharply the curve bends; that number is the curvature of the curve.

Figure 11.5.3 : Establishing the concept of curvature.


We derive this number in the following way. Consider Figure 11.5.3b, where unit tangent vectors are graphed around points A and
B . Notice how the direction of the unit tangent vector changes quite a bit near A , whereas it does not change as much around B .

This leads to an important concept: measuring the rate of change of the unit tangent vector with respect to arc length gives us a
measurement of curvature.

Definition 11.5.1: Curvature

Let ⇀
r (s) be a vector-valued function where s is the arc length parameter. The curvature κ of the graph of ⇀
r (s) is

dT ⇀

κ =∥ ∥ = ∥ T (s)∥. (11.5.5)
ds

If ⇀
r (s) is parametrized by the arc length parameter, then

⇀′ ′
⇀ r (s) ⇀ T (s)
T(s) = and N(s) = . (11.5.6)
⇀′ ⇀

∥ r (s)∥ ∥ T (s)∥


Having defined ∥T ′
(s)∥ = κ , we can rewrite the second equation as
⇀ ⇀

T (s) = κ N(s). (11.5.7)

11.5.3 https://math.libretexts.org/@go/page/4224
⇀ ⇀ ⇀
We already knew that T ′
(s) is in the same direction as N(s) ; that is, we can think of T(s) as being "pulled'' in the direction of
⇀ ⇀
N(s). How "hard'' is it being pulled? By a factor of κ . When the curvature is large, T(s) is being "pulled hard'' and the direction of

T(s) changes rapidly. When κ is small, T (s) is not being pulled hard and hence its direction is not changing rapidly.

We use Definition 11.5.1 to find the curvature of the line in Example 11.5.2.

Example 11.5.2: Finding the curvature of a line

Use Definition 11.5.1 to find the curvature of ⇀


r (t) = ⟨3t − 1, 4t + 2⟩ .
Solution
In Example ??? , we found that the arc length parameter was defined by s = 5t , so ⇀
r (s) = ⟨3t/5 − 1, 4t/5 + 2⟩ parametrized

r with the arc length parameter. To find κ , we need to find T (s) .
⇀ ′


⇀′
T(s) = r (s) (recall this is a unit vector)

= ⟨3/5, 4/5⟩.

Therefore


T (s) = ⟨0, 0⟩

and


κ = ∥ T (s)∥ = 0.

It probably comes as no surprise that the curvature of a line is 0. (How "curvy'' is a line? It is not curvy at all.)

While the definition of curvature is a beautiful mathematical concept, it is nearly impossible to use most of the time; writing r in ⇀

terms of the arc length parameter is generally very hard. Fortunately, there are other methods of calculating this value that are much
easier. There is a tradeoff: the definition is "easy'' to understand though hard to compute, whereas these other formulas are easy to
compute though it may be hard to understand why they work.

theorem 100: Formulas for Curvature

Let C be a smooth curve on an open interval I in the plane or in space.


1. If C is defined by y = f (x), then
′′
|f (x)|
κ = . (11.5.8)
3/2
2

(1 + (f (x)) )

2. If C is defined as a vector-valued function in the plane, ⇀


r (t) = ⟨x(t), y(t)⟩ , then
′ ′′ ′′ ′
|x y −x y |
κ = . (11.5.9)
3/2
′ 2 ′ 2
((x ) + (y ) )

3. If C is defined in space by a vector-valued function ⇀


r (t) , then
⇀ ⇀
′ ⇀′ ⇀′′ ⇀
∥ T (t)∥ ∥ r (t) × r (t)∥ a (t) ⋅ N(t)
κ = = = . (11.5.10)
⇀′ ⇀′ ⇀ 2
3 ∥ v (t)∥
∥ r (t)∥ ∥ r (t)∥

We practice using these formulas.

Example 11.5.3: Finding the curvature of a circle

Find the curvature of a circle with radius r, defined by ⇀


c (t) = ⟨r cos t, r sin t⟩ .

Solution

11.5.4 https://math.libretexts.org/@go/page/4224
Before we start, we should expect the curvature of a circle to be constant, and not dependent on t . (Why?)
We compute κ using the second part of Theorem 100.
|(−r sin t)(−r sin t) − (−r cos t)(r cos t)|
κ =
3/2
2 2
((−r sin t) + (r cos t) )

2 2 2
r (sin t + cos t)
=
2 3/2
2 2
(r (sin t + cos t))

2
r 1
= = .
3
r r

We have found that a circle with radius r has curvature κ = 1/r .

Example 11.5.3 gives a great result. Before this example, if we were told "The curve has a curvature of 5 at point A ,'' we would
have no idea what this really meant. Is 5 "big'' - does is correspond to a really sharp turn, or a not-so-sharp turn? Now we can think
of 5 in terms of a circle with radius 1/5. Knowing the units (inches vs. miles, for instance) allows us to determine how sharply the
curve is curving.
Let a point P on a smooth curve C be given, and let κ be the curvature of the curve at P . A circle that:
passes through P ,
lies on the concave side of C ,
has a common tangent line as C at P and
has radius r = 1/κ (hence has curvature κ )
is the osculating circle, or circle of curvature, to C at P , and r is the radius of curvature. Figure 11.5.4 shows the graph of the
curve seen earlier in Figure 11.30 and its osculating circles at A and B . A sharp turn corresponds to a circle with a small radius; a
gradual turn corresponds to a circle with a large radius. Being able to think of curvature in terms of the radius of a circle is very
useful.

Figure 11.30

Example 11.5.4: Finding curvature

Find the curvature of the parabola defined by y = x at the vertex and at x = 1 .


2

Solution
We use the first formula found in Theorem 100.
|2|
κ(x) =
3/2
2
(1 + (2x) )

2
= .
3/2
2
(1 + 4x )

11.5.5 https://math.libretexts.org/@go/page/4224
Figure 11.5.5 : Examining the curvature of y = x . 2

At the vertex (x = 0 ), the curvature is κ = 2 . At x = 1 , the curvature is κ = 2/(5) ≈ 0.179. So at x = 0 , the curvature of
3/2

y =x
2
is that of a circle of radius 1/2; at x = 1 , the curvature is that of a circle with radius ≈ 1/0.179 ≈ 5.59. This is
illustrated in Figure 11.5.5. At x = 3 , the curvature is 0.009; the graph is nearly straight as the curvature is very close to 0.

Example 11.5.5: Finding curvature

Find where the curvature of ⇀ 2 3


r (t) = ⟨t, t , 2 t ⟩ is maximized.

Solution
We use the third formula in Theorem 100 as ⇀
r (t) is defined in space. We leave it to the reader to verify that
⇀′ 2 ⇀′′ ⇀′ ⇀′′ 2
r (t) = ⟨1, 2t, 6 t ⟩, r (t) = ⟨0, 2, 12t⟩, and r (t) × r (t) = ⟨12 t , −12t, 2⟩. (11.5.11)

Figure 11.5.6 : Understanding the curvature of a curve in space.


Thus
⇀′ ⇀′′
∥ r (t) × r (t)∥
κ(t) =
⇀′ 3
∥ r (t)∥

2
∥⟨12 t , −12t, 2⟩∥
=
2 3
∥⟨1, 2t, 6 t ⟩∥

− −−−−−−−−−−−− −
√ 144 t4 + 144 t2 + 4
=
− −−−− − −−−−− − 3
2 4
(√ 1 + 4 t + 36 t )

While this is not a particularly "nice'' formula, it does explicitly tell us what the curvature is at a given t value. To maximize
κ(t) , we should solve κ (t) = 0 for t . This is doable, but very time consuming. Instead, consider the graph of κ(t) as given in

Figure 11.5.6a. We see that κ is maximized at two t values; using a numerical solver, we find these values are t ≈ ±0.189. In
Figure 11.5.1b we graph r (t) and indicate the points where curvature is maximized.

11.5.6 https://math.libretexts.org/@go/page/4224
Curvature and Motion
Let ⇀
r (t) be a position function of an object, with velocity ⇀
v (t) = r (t)
⇀′
and acceleration ⇀
a (t) = r
⇀′′
(t) . In Section 11.4 we
⇀ ⇀
established that acceleration is in the plane formed by T(t) and N(t), and that we can find scalars a and a such that T N

⇀ ⇀

a (t) = aT T(t) + aN N(t). (11.5.12)

Theorem 98 gives formulas for a and a :T N

⇀ ⇀
d ∥ v (t) × a (t)∥

aT = (∥ v (t)∥) and aN = . (11.5.13)

dt ∥ v (t)∥


We understood that the amount of acceleration in the direction of T relates only to how the speed of the object is changing, and

that the amount of acceleration in the direction of N relates to how the direction of travel of the object is changing. (That is, if the
object travels at constant speed, a = 0 ; if the object travels in a constant direction, a = 0 .)
T N

In Equation 11.5.3 at the beginning of this section, we found s ′ ⇀


(t) = ∥ v (t)∥ . We can combine this fact with the above formula for
aT to write
d d
⇀ ′ ′′
aT = (∥ v (t)∥) = (s (t)) = s (t). (11.5.14)
dt dt

Since s (t) is speed, s (t) is the rate at which speed is changing with respect to time. We see once more that the component of
′ ′′

acceleration in the direction of travel relates only to speed, not to a change in direction.
Now compare the formula for a above to the formula for curvature in Theorem 100:
N

⇀ ⇀ ⇀′ ⇀′′ ⇀ ⇀
∥ v (t) × a (t)∥ ∥ r (t) × r (t)∥ ∥ v (t) × a (t)∥
aN = and κ = = . (11.5.15)
⇀ ⇀′ ⇀ 3
∥ v (t)∥ 3 ∥ v (t)∥
∥ r (t)∥

Thus
⇀ 2
aN = κ∥ v (t)∥ (11.5.16)

2

= κ(s (t))

This last equation shows that the component of acceleration that changes the object's direction is dependent on two things: the
curvature of the path and the speed of the object.
Imagine driving a car in a clockwise circle. You will naturally feel a force pushing you towards the door (more accurately, the door
is pushing you as the car is turning and you want to travel in a straight line). If you keep the radius of the circle constant but speed
up (i.e., increasing s (t) ), the door pushes harder against you (a has increased). If you keep your speed constant but tighten the

N

turn (i.e., increase κ ), once again the door will push harder against you.
Putting our new formulas for a and a together, we have
T N

⇀ ⇀
⇀ ′′ ⇀ 2
a (t) = s (t)T(t) + κ∥ v (t)∥ N(t). (11.5.17)

This is not a particularly practical way of finding aT and a , but it reveals some great concepts about how acceleration interacts
N

with speed and the shape of a curve.

Example 11.5.6: Curvature and road design

The minimum radius of the curve in a highway cloverleaf is determined by the operating speed, as given in the table in Figure
11.5.7. For each curve and speed, compute a . N

Figure 11.5.7 : Operating speed and minimum radius in highway cloverleaf design.
Solution

11.5.7 https://math.libretexts.org/@go/page/4224
Using Equation 11.5.16, we can compute the acceleration normal to the curve in each case. We start by converting each speed
from "miles per hour'' to "feet per second'' by multiplying by 5280/3600.
35mph, 310ft ⇒ 51.33ft/s, κ = 1/310

⇀ 2
aN = κ ∥ v (t)∥

1 2
= (51.33 )
310

2
= 8.50 ft/s .

40mph, 430ft ⇒ 58.67ft/s, κ = 1/430

1 2
aN = (58.67 )
430

2
= 8.00 ft/s .

45mph,540ft ⇒ 66ft/s, κ = 1/540

1 2
aN = (66 )
540

2
= 8.07 ft/s .

Note that each acceleration is similar; this is by design. Considering the classic "Force = mass × acceleration'' formula, this
acceleration must be kept small in order for the tires of a vehicle to keep a "grip'' on the road. If one travels on a turn of radius
310ft at a rate of 50mph, the acceleration is double, at 17.35ft/s . If the acceleration is too high, the frictional force created by
2

the tires may not be enough to keep the car from sliding. Civil engineers routinely compute a "safe'' design speed, then subtract
5-10mph to create the posted speed limit for additional safety.

We end this chapter with a reflection on what we've covered. We started with vector-valued functions, which may have seemed at
the time to be just another way of writing parametric equations. However, we have seen that the vector perspective has given us
great insight into the behavior of functions and the study of motion. Vector-valued position functions convey displacement,
distance traveled, speed, velocity, acceleration and curvature information, each of which has great importance in science and
engineering.

This page titled 11.5: The Arc Length Parameter and Curvature is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or
curated by Gregory Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history
is available upon request.

11.5.8 https://math.libretexts.org/@go/page/4224
11.E: Applications of Vector Valued Functions (Exercises)
11.1: Vector-Valued Functions
Terms and Concepts
1. Vector-valued functions are closely related to _______ ________ of graphs.
2. When sketching vector-valued functions, technically one isn't graphing points, but rather _______.
3. It can be useful to think of _______ as a vector that points from a starting position to an ending position.

Problems
In Exercises 4-11, sketch the vector-valued function on the given interval.
4. r (t)
⃗ = ⟨t
2
,t
2
− 1⟩, for − 2 ≤ t ≤ 2 .
5. r (t)
⃗ = ⟨t
2 3
, t ⟩, for − 2 ≤ t ≤ 2 .
6. r (t)
⃗ = ⟨1/t, t/ t
2
⟩, for − 2 ≤ t ≤ 2 .
7. r (t)
⃗ =⟨
1

10
2
t , sin t⟩, for − 2π ≤ t ≤ 2π .
8. r (t)
⃗ =⟨
1

10
2
t , sin t⟩, for − 2π ≤ t ≤ 2π .
9. r (t)
⃗ = ⟨3 sin(πt), 2 cos(πt)⟩, on [0, 2] .
10. r (t)
⃗ = ⟨3 cos t, 2 sin(2t)⟩, on [0, 2π] .
11. r (t)
⃗ = ⟨2 sec t, tan t⟩, on [−π, π] .
In Exercises 12-15, sketch the vector-valued function on the given interval in R . Technology may be useful in creating the
3

sketch.
12. r (t)
⃗ = ⟨2 cos t, t, 2 sin t⟩, on [0, 2π] .
13. r (t)
⃗ = ⟨3 cos t, sin t, t/π⟩, on [0, 2π] .
14. r (t)
⃗ = ⟨cos t, sin t, sin t⟩, on [0, 2π] .
15. r (t)
⃗ = ⟨cos t, sin t, sin(2t)⟩, on [0, 2π] .
In Exercises 16-19, find ∥r (t)∥
⃗ .
16. r (t)
⃗ = ⟨t, t
2
⟩ .
17. r (t)
⃗ = ⟨5 cos t, 3 sin t⟩ .

18. r (t)
⃗ = ⟨2 cos t, 2 sin t, t⟩ .

19. r (t)
⃗ = ⟨cos t, t, t ⟩ .
2

In Exercises 20-27, create a vector-valued function whose graph matches the given description.
20. A circle of radius 2, centered (1, 2), traced counter-clockwise once on [0, 2π].
21. A circle of radius 3, centered (5, 5), traced clockwise once on [0, 2π].
22. An ellipse, centered at (0, 0) with vertical major axis length 10 and minor axis of length 3, traced once counter-clockwise on
[0, 2π].

23. An ellipse, centered at (3, −2) with horizontal major axis length 6 and minor axis of length 4, traced once clockwise on [0, 2π].
24. A line through (2, 3) with a slope of 5.
25. A line through (1, 5) with a slope of -1/2.
26. A vertically oriented helix with radius of 2 that starts at (2, 0, 0) and ends at (2, 0, 4π) after 1 revolution on [0, 2π].
27. A vertically oriented helix with radius of 3 that starts at (3, 0, 0) and ends at (3, 0, 3) after 2 revolutions on [0, 1].

11.E.1 https://math.libretexts.org/@go/page/10002
In Exercises 28-31, find the average rate of change of r (t)
⃗ on the given interval.
28. r (t)
⃗ = ⟨t, t
2
⟩ on [−2, 2].

29. r (t)
⃗ = ⟨t, t + sin t⟩ on [0, 2π].

30. r (t)
⃗ = ⟨3 cos t, 2 sin t, t⟩ on [0, 2π].

31. r (t)
⃗ = ⟨t, t
2 3
, t ⟩ on [−1, 3].

11.2: Calculus and Vector-Valued Functions


Terms and Concepts
1. Limits, derivatives and integrals of vector-valued functions are all evaluated _______ -wise.
2. The definite integral of a rate of change function gives ________.
3. Why is it generally not useful to graph both r (t)
⃗ and r ⃗ (t) on the same axes?

Problems
In Exercises 4-7, evaluate the given limit.
4. lim⟨2t + 1, 3t 2
− 1, sin t⟩
t→5

5. lim⟨e t
,
t −9

t+3

t→3

6. lim⟨ t

sin t
, (1 + t) t ⟩
t→0


r (t+h)− ⃗
r (t)
7. lim h

, where r (t) = ⟨t , t, 1⟩
2

h→0

In Exercises 8-9, identify the interval(s) on which r (t)


⃗ is continuous.
8. r (t)
⃗ = ⟨t
2
, 1/t⟩

9. r (t)
⃗ = ⟨cos t, e
t
, ln t⟩

In Exercises 10-14, find the derivative of the given function.


10. r (t)
⃗ = ⟨cos t, e
t
, ln t⟩

11. r (t)
⃗ =⟨
1

t
,
2t−1

3t+1
, tan t⟩

12. r (t)
⃗ = (t
2
)⟨sin t, 2t + 5⟩

13. r (t)
⃗ = ⟨t
2
+ 1, t − 1⟩ ⋅ ⟨sin t, 2t + 5⟩

14. r (t)
⃗ = ⟨t
2
+ 1, t − 1, 1⟩ × ⟨sin t, 2t + 5, 1⟩

′ ′ ′
In Exercises 15-18, find r ⃗ (t) . Sketch r (t)
⃗ and r ⃗ (1) , with the initial point of r ⃗ (1) at r (1)
⃗ .
15. r (t)
⃗ = ⟨t
2
+ t, t
2
− t⟩

16. r (t)
⃗ = ⟨t
2
− 2t + 2, t
3
− 3t
2
+ 2t⟩

17. r (t)
⃗ = ⟨t
2
+ 1, t
3
− t⟩

18. r (t)
⃗ = ⟨t
2
− 4t + 5, t
3
− 6t
2
+ 11t − 6⟩

In Exercises 19-22, give the equation of the line tangent to the graph of r (t)
⃗ at the given t value.
19. r (t)
⃗ = ⟨t
2
+ t, t
2
− t⟩ at t = 1.

20. r (t)
⃗ = ⟨3 cos t, sin t⟩ at t = π/4.

21. r (t)
⃗ = ⟨3 cos t, 3 sin t, t⟩ at t = π.

22. r (t)
⃗ = ⟨e
t
, tan t, t⟩ at t = 0.

In Exercises 23-26, find the value(s) of t for which r (t)


⃗ is not smooth.

11.E.2 https://math.libretexts.org/@go/page/10002
23. r (t)
⃗ = ⟨cos t, sin t − t⟩

24. r (t)
⃗ = ⟨t
2
− 2t + 1, t
3
+t
2
− 5t + 3⟩

25. r (t)
⃗ = ⟨cos t − sin t − cos t, cos(4t)⟩

26. r (t)
⃗ = ⟨t
3
− 3t + 2, − cos(πt), sin (πt)⟩
2

Exercises 27-29 ask you to verify parts of Theorem 92. In each let ⃗
f (t) = t , r (t)
3 2
= ⟨t , t − 1, 1⟩ and ⃗
s (t)
t
= ⟨sin t, e , t⟩ .
Compute the various derivatives as indicated.
′ ′
27. Simplify f (t)r (t)
⃗ , then find its derivative; show this is the same as f ′
(t)r ⃗ (t) + f (t)r ⃗ (t) .
′ ′ ′
28. Simplify r (t)
⃗ ⃗
⋅ s (t) , then find its derivative; show this is the same as r ⃗ (t) ⋅ s ⃗ (t) + r (t)
⃗ ⋅ s ⃗ (t) .

′ ′ ′
29. Simplify r (t)
⃗ ⃗
× s (t) , then find its derivative; show this is the same as r ⃗ (t) × s ⃗ (t) + r (t)
⃗ × s ⃗ (t) .

In Exercises 30-33, evaluate the given definite or indefinite integral.


30. ∫ ⟨t 3
, cos t, te ⟩ dt
t

31. ∫ ⟨ 1
2
, sec
2
t⟩ dt
1+t

π
32. ∫0
⟨− sin t, cos t⟩ dt

2
33. ∫−2
⟨2t + 1, 2t − 1⟩ dt

In Exercises 34-37, solve the given initial value problems.



34. Find r (t)
⃗ , given that r ⃗ (t) = ⟨t, sin t⟩ and r (0)
⃗ = ⟨2, 2⟩.


35. Find r (t)
⃗ , given that r ⃗ (t) = ⟨1, (t + 1), tan t⟩ and r (0)
⃗ = ⟨1, 2⟩.

′′ ′
36. Find r (t)
⃗ , given that r ⃗ 2
(t) = ⟨t , t, 1⟩ , r ⃗ (0) = ⟨1, 2, 3⟩ and r (0)
⃗ = ⟨4, 5, 6⟩ .

′′ ′
37. Find r (t)
⃗ , given that r ⃗ (t) = ⟨cos t, sin t, e ⟩
t
, r ⃗ (0) = ⟨0, 0, 0⟩ and r (0)
⃗ = ⟨0, 0, 0⟩ .

In Exercises 38-41, find the arc length of r (t)


⃗ on the indicated interval.
38. r (t)
⃗ = ⟨2 cos t, 2 sin t, 3t⟩ on [0, 2π]

39. r (t)
⃗ = ⟨5 cos t, 3 sin t, 4 sin t⟩ on [0, 2π] .

40. r (t)
⃗ = ⟨t
3 2 3
, t , t ⟩ on [0, 1] .
41. r (t)
⃗ = ⟨e
−t
cos t, e
−t
sin t⟩ on [0, 1] .

42. Prove Theorem 93; that is, show if r (t)
⃗ has constant length and is differentiable, then ⃗
r (t) ⋅ r ⃗ (t) = 0 . (Hint: use the Product
Rule to compute (r (t)
⃗ ⃗
⋅ r (t)) .) d

dt

11.3: The Calculus of Motion


Terms and Concepts
1. How is velocity different from speed?
2. What is the difference between displacement and distance traveled?
3. What is the difference between average velocity and average speed?
4. Distance traveled is the same as ______ _______, just viewed in a different context.
5. Describe a scenario where an object's average speed is a large number, but the magnitude of the average velocity is not a large
number.
6. Explain why it is not possible to have an average velocity with a large magnitude but a small average speed.

Problems
In Exercises 7-10, a position function r (t)
⃗ is given. Find v(t)
⃗ and a⃗(t) .
7. r (t)
⃗ = ⟨2t + 1, 5t − 2, 7⟩

11.E.3 https://math.libretexts.org/@go/page/10002
8. r (t)
⃗ = ⟨3 t
2
− 2t + 1, −t
2
+ t + 14⟩

9. r (t)
⃗ = ⟨cos t, sin t⟩

10. r (t)
⃗ = ⟨t/10, − cos t, sin t⟩

In Exercises 11-14, a position function r (t) ⃗ is given. Sketch ⃗


r (t) and a⃗(t) , then add ⃗ t0 )
r( and a⃗(t0 ) to your sketch, with
their initial points at r (t ) , for the given value of t .
⃗ 0 0

11. r (t)
⃗ = ⟨t, sin t⟩ on [0, π/2]; t0 = π/4

−−−
12. r (t)
⃗ = ⟨t
2 2
, sin t ⟩ on [0, π/2]; t0 = √π/4

13. r (t)
⃗ = ⟨t
2
+ t, −t
2
+ 2t⟩ on [−2, 2]; t0 = 1

2t+3
14. r (t)
⃗ =⟨ 2
t +1
2
, t ⟩ on [−1, 1]; t0 = 0

In Exercises 15-24, a position function r (t)


⃗ of an object is given. Find the speed of the object in terms of t , and find where
the speed is minimized/maximized on the indicated interval.
15. r (t)
⃗ = ⟨t
2
, t⟩ on [−1, 1]

16. r (t)
⃗ = ⟨t
2
,t
2 3
− t ⟩ on [−1, 1]

17. r (t)
⃗ = ⟨5 cos t, 5 sin t⟩ on [0, 2π]

18. r (t)
⃗ = ⟨2 cos t, 5 sin t⟩ on [0, 2π]

19. r (t)
⃗ = ⟨sec t, tan t⟩ on [0, π/4]

20. r (t)
⃗ = ⟨t + cos t, 1 − sin t⟩ on [0, 2π]

21. r (t)
⃗ = ⟨12t, 5 cos t, 5 sin t⟩ on [0, 4π]

22. r (t)
⃗ = ⟨t
2
− t, t
2
+ t, t⟩ on [0, 1]

−−−− −
23. ⃗
r (t)
2 2
= ⟨t, t , √1 − t ⟩ on [−1, 1]

2 v0 sin θ
24. Projectile Motion: r (t)
⃗ = ⟨(v 0 cos θ)t, −
1

2
2
gt + (v0 sin θ)t⟩ on [0,
g
] .

In Exercises 25-28, position functions r ⃗ (t) and r ⃗ (s) for two objects are given that follow the same path on the respective
1 2

intervals.
(a) Show that the positions are the same at the indicated t and s values; ie., show r ⃗ (t ) = r ⃗ (s ) .
0 0 1 0 2 0

(b) Find the velocity, speed and acceleration of the two objects at t and s , respectively. 0 0

25.
2
⃗ (t) = ⟨t, t ⟩ on [0, 1]; t0 = 1
r1
2 4
⃗ (t) = ⟨s , s ⟩ on [0, 1]; s0 = 1
r2

26.
⃗ (t) = ⟨3 cos t, 3 sin t⟩ on [0, 2π]; t0 = π/2
r1

⃗ (t) = ⟨3 cos(4s), 3 sin(4s)⟩ on [0, π/2]; s0 = π/8


r2

27.
⃗ (t) = ⟨3t, 2t⟩ on [0, 2]; t0 = 2
r1

⃗ (t) = ⟨6t − 6, 4t − 4⟩ on [1, 2]; s0 = 2


r2

28.
⃗ (t) = ⟨t, √t⟩ on [0, 1]; t0 = 1
r1
−−−−
r2⃗ (t) = ⟨sin t, √sin t ⟩ on [0, π/2]; s0 = π/2

In Exercises 29-32, find the position function of an object given its acceleration and initial velocity and position.
29. a⃗(t) = ⟨2, 3⟩; v⃗(0) = ⟨1, 2⟩, r (0)
⃗ = ⟨5, −2⟩

30. a⃗(t) = ⟨2, 3⟩; ⃗


v(1) = ⟨1, 2⟩, ⃗
r (1) = ⟨5, −2⟩

31. a⃗(t) = ⟨cos t, − sin t⟩; ⃗


v(0) = ⟨0, 1⟩, ⃗
r (0) = ⟨0, 0⟩

11.E.4 https://math.libretexts.org/@go/page/10002
32. a⃗(t) = ⟨0, −32⟩; ⃗
v(0) = ⟨10, 50⟩, ⃗
r (0) = ⟨0, 0⟩

In Exercises 33-36, find the displacement, distance traveled, average velocity, and average speed of the described object on
the given interval.
33. An object with position function ⃗
r (t) = ⟨2 cos t, 2 sin t, 3t⟩ , where distances are measured in feet and time is in seconds, on
[0, 2π].

34. An object with position function ⃗


r (t) = ⟨5 cos t, −5 sin t⟩ , where distances are measured in feet and time is in seconds, on
[0, π].

35. An object with velocity function v(t)


⃗ = ⟨cos t, sin t⟩ , where distances are measured in feet and time is in seconds, on [0, 2π].

36. An object with velocity function r (t)


⃗ = ⟨1, 2, −1⟩, where distances are measured in feet and time is in seconds, on [0, 10].

Exercises 37-42 ask you to solve a variety of problems based on the principles of projectile motion.
37. A boy whirls a ball, attached to a 3ft string, above his head in a counter–clockwise circle. The ball makes 2 revolutions per
second.
At what t-values should the boy release the string so that the ball heads directly for a tree standing 10ft in front of him?
38. David faces Goliath with only a stone in a 3ft sling, which he whirls above his head at 4 revolutions per second. They stand 20ft
apart.
(a) At what t-values must David release the stone in his sling in order to hit Goliath?
(b) What is the speed at which the stone is traveling when released?
(c) Assume David releases the stone from a height of 6ft and Goliath's forehead is 9ft above the ground. What angle of elevation
must David apply to the stone to hit Goliath's head?
39. A hunter aims at a deer which is 40 yards away. Her cross-bow is at a height of 5ft, and she aims for a spot on the deer 4ft
above the ground. The crossbow fires her arrows at 300ft/s.
(a) At what angle of elevation should she hold the crossbow to hit her target?
(b) If the deer is moving perpendicularly to her line of sight at a rate of 20mph, by approximately how much should she lead the
deer in order to hit in the desired location?
40. A baseball player hits a ball at 100mph, with an intial height of 3ft and an angle of elevation of 20 , at Boston's Fenway Park.

The ball flies towards the famed "Green Monster," a wall 37ft high located 310ft from home plate.
(a) Show that as hit, the ball hits the wall.
(b) Show that if the angle of elevation is 21 , the ball clears the Green Monster.

41. A Cessna flies at 1000ft at 150mph and drops a box of supplies to the professor (and his wife) on an island. Ignoring wind
resistance, how far horizontally will the supplies travel before they land?
42. A football quarterback throws a pass from a height of 6ft, intending to hit his receiver 20yds away at a height of 5ft.
(a) If the ball is thrown at a rate of 50mph, what angle of elevation is needed to hit his intended target?
(b) If the ball is thrown at with an angle of elevation of 8 , what initial ball speed is needed to hit his target?

11.4: Unit Tangent and Normal Vectors


Terms and Concepts
1. If T (t)

is a unit tangent vector, what is ∥T (t)∥

?

2. If N ⃗ (t) is a unit normal vector, what is N ⃗ (t) ⋅ r ⃗ (t) ?
3. The acceleration vector a⃗(t) lies in the plane defined by what two vectors?
4. a measures how much the acceleration is affecting the _______ of an object.
T

Problems
In Exercises 5-8, given r (t)
⃗ , find T (t)

and evaluate it at the indicated value of t.
5. r (t)
⃗ = ⟨2 t
2 2
,t − 1⟩, t =1

6. r (t)
⃗ = ⟨t, cos t⟩, t = π/4

11.E.5 https://math.libretexts.org/@go/page/10002
7. r (t)
⃗ = ⟨cos
3
t, sin
3
t⟩, t = π/4

8. r (t)
⃗ = ⟨cos t, sin t⟩, t =π

In Exercises 9-12, find the equation of the line tangent to the curve at the indicated t-value using the unit tangent vector.
Note: these are the same problems as in Exercises 5-8.
9. r (t)

2
= ⟨2 t ,t
2
− 1⟩, t =1

10. r (t)
⃗ = ⟨t, cos t⟩, t = π/4

11. r (t)
⃗ = ⟨cos
3
t, sin
3
t⟩, t = π/4

12. r (t)
⃗ = ⟨cos t, sin t⟩, t =π

In Exercises 13-16, find N ⃗ (t) using Definition 75. Confirm the result using the Theorem 97.
13. r (t)
⃗ = ⟨3 cos t, 3 sin t⟩

14. r (t)
⃗ = ⟨t, t
2

15. r (t)
⃗ = ⟨cos t, 2 sin t⟩

16. r (t)
⃗ = ⟨e
t
,e
−t

In Exercises 17-20, a position function r (t)


⃗ is given along with its unit tangent vector ⃗
T (t) evaluated at t =a , for some
value of a .
(a) Confirm that T (a)

is as stated.
(b) Using a graph of r (t)
⃗ and Theorem 97, find N ⃗ (a) .

17. r (t)
⃗ = ⟨3 cos t, 5 sin t⟩;

T (π/4) = ⟨−
√34
3
,
√34
5

18. r (t)
⃗ = ⟨t,
2
1
⟩;

T (1) = ⟨
2
,−
1

t +1 √5 √5

19. r (t)
⃗ = (1 + 2 sin t)(cos t, sin t);

T (0) = ⟨
2
,
1

√5 √5

20. r (t)
⃗ = ⟨cos
3
t, sin
3
t⟩ ;

T (π/4) = ⟨−
1

√2
,−
1

√2

In Exercises 21-24, find N ⃗ (t).


21. r (t)
⃗ = ⟨4t, 2 sin t, 2 cos t⟩

22. r (t)
⃗ = ⟨5 cos t, 3 sin t, 4 sin t⟩

23. r (t)
⃗ = ⟨a cos t, a sin t, bt⟩

24. r (t)
⃗ = ⟨cos(at), sin(at), t⟩

In Exercises 25-30, find a and a T N given r (t)


⃗ . Sketch r (t)
⃗ on the indicated interval, and comment on the relative sizes of a T

and a at the indicated t values.


N

25. r (t)
⃗ = ⟨t, t
2
⟩ on [−1, 1]; consider t = 0 and t = 1 .
26. r (t)
⃗ = ⟨t, 1/t⟩ on (0, 4; consider t = 1 and t = 2 .
27. r (t)
⃗ = ⟨2 cos t, 2 sin t⟩ on [0, 2π]; consider t = 1 and t = π/2 .
−−− −
28. ⃗
r (t)
2 2
= ⟨cos(t ), sin(t )⟩ on (0, 2π]; consider t = √π/2 and t = √π .
29. r (t)
⃗ = ⟨a cos t, a sin t, bt⟩ on [0, 2π]; where a, b > 0; consider t = 0 and t = π/2 .
30. r (t)
⃗ = ⟨5 cos t, 4 sin t, 3 sin t⟩ on [0, 2π]; consider t = 0 and t = π/2 .

11.5: The Arc Length Parameter and Curvature

11.E.6 https://math.libretexts.org/@go/page/10002
Terms and Concepts
1. It is common to describe position in terms of both ______ and/or _______.
2. A measure of the "curviness" of a curve is ________.
3. Give two shapes with constant curvature.
4. Describe in your own words what an "osculating circle" is.

5. Complete the identity: T ⃗ (s) = _________N ⃗ (s) .


6. Given a position function r (t)
⃗ , how are a and a T N affected by the curvature?

Problems
In Exercises 7-10, a position function r (t)
⃗ is given, where \(t=0\) corresponds to the initial position. Find the arc length
parameter s , and rewrite r (t)
⃗ in terms of s ; that is, find r (s)
⃗ .
7. r (t)
⃗ = ⟨2t, t, −2t⟩

8. r (t)
⃗ = ⟨7 cos t, 7 sin t⟩

9. r (t)
⃗ = ⟨3 cos t, 3 sin t, 2t⟩

10. r (t)
⃗ = ⟨5 cos t, 13 sin t, 12 cos t⟩

In Exercises 11-22, a curve C is described along with 2 points on C.


(a) Using a sketch, determine at which of these points the curvature is greater.
(b) Find the curvature κ of C, and evaluate κ at each of the 2 given points.
11. C is defined by y = x 3
−x ; points given at x = 0 and x = 1/2.
12. C is defined by y = 2
1

x +1
; points given at x = 0 and x = 2 .

13. C is defined by y = cos x ; points given at x = 0 and x = π/2.


−−−−−
14. C is defined by y = √1 − x on (−1, 1); points given at x = 0 and x = 1/2.
2

15. C is defined by r (t)


⃗ = ⟨cos t, sin(2t)⟩ ; points given at t = 0 and t = π/4 .

16. C is defined by r (t)


⃗ = ⟨cos
2
(t), sin t cos t⟩ ; points given at t = 0 and t = π/3 .
17. C is defined by r (t)
⃗ = ⟨t
2
− 1, t
3
− t⟩ ; points given at t = 0 and t = π/6 .
18. C is defined by r (t)
⃗ = ⟨tan t, sec t⟩ ; points given at t = 0 and t = π/6 .

19. C is defined by r (t)


⃗ = ⟨4t + 2, 3t − 1, 2t + 5⟩ ; points given at t = 0 and t = 1 .

20. C is defined by r (t)


⃗ = ⟨t
3
− t, t
3
− 4, t
2
− 1⟩ ; points given at t = 0 and t = 1 .
21. C is defined by r (t)
⃗ = ⟨3 cos t, 3 sin t, 2t⟩ ; points given at t = 0 and t = π/2 .

22. C is defined by r (t)


⃗ = ⟨5 cos t, 13 sin t, 12 cos t⟩ ; points given at t = 0 and t = π/2 .

In Exercises 23-26, find the value of x or t where curvature is maximized.


23. y = 1

6
x
3

24. y = sin x
25. r (t)
⃗ = ⟨t
2 2
+ 2t, 3t − t ⟩

26. r (t)
⃗ = ⟨t, 4/t, 3/t⟩

In Exercises 27-30, find the radius of curvature at the indicated value.


27. y = tan x, at x = π/4

28. y = x 2
+ x − 3, at x = π/4

29. r (t)
⃗ = ⟨cos t, sin(3t)⟩, at t = 0

30. r (t)
⃗ = ⟨5 cos(3t), t⟩, at t = 0

11.E.7 https://math.libretexts.org/@go/page/10002
In Exercises 31-34, find the equation of the osculating circle to the curve at the indicated t-value.
31. r (t)
⃗ = ⟨t, t
2
⟩, at t = 0.

32. r (t)
⃗ = ⟨3 cos t, sin t⟩, at t = 0.

33. r (t)
⃗ = ⟨3 cos t, sin t⟩, at t = π/2.

34. r (t)
⃗ = ⟨t
2
− t, t
2
+ t⟩, at t = 0.

11.E: Applications of Vector Valued Functions (Exercises) is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by
LibreTexts.

11.E.8 https://math.libretexts.org/@go/page/10002
CHAPTER OVERVIEW
12: Functions of Several Variables
A function of the form y = f (x) is a function of a single variable; given a value of x , we can find a value y . Even the vector--
valued functions of Chapter 11 are single--variable functions; the input is a single variable though the output is a vector. There are
many situations where a desired quantity is a function of two or more variables. For instance, wind chill is measured by knowing
the temperature and wind speed; the volume of a gas can be computed knowing the pressure and temperature of the gas; to compute
a baseball player's batting average, one needs to know the number of hits and the number of at--bats. This chapter studies
multivariable functions, that is, functions with more than one input.
12.1: Introduction to Multivariable Functions
12.2: Limits and Continuity of Multivariable Functions
12.3: Partial Derivatives
12.4: Differentiability and the Total Differential
12.5: The Multivariable Chain Rule
12.6: Directional Derivatives
12.7: Tangent Lines, Normal Lines, and Tangent Planes
12.8: Extreme Values
12.E: Applications of Functions of Several Variables (Exercises)

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 12: Functions of Several Variables is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by
Gregory Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available
upon request.

1
12.1: Introduction to Multivariable Functions
Definition 77 Function of Two Variables
Let D be a subset of R . A function f of two variables is a rule that assigns each pair (x, y) in D a value z = f (x, y) in R.
2

D is the domain of f ; the set of all outputs of f is the range.

Example 12.1.1: Understanding a function of two variables

Let z = f (x, y) = x 2
−y . Evaluate f (1, 2), f (2, 1), and f (−2, 4); find the domain and range of f .
Solution
Using the definition f (x, y) = x 2
−y , we have:
2
f (1, 2) = 1 − 2 = −1
2
f (2, 1) = 2 −1 = 3
2
f (−2, 4) = (−2 ) −4 = 0

The domain is not specified, so we take it to be all possible pairs in R for which f is defined. In this example, f is defined for
2

all pairs (x, y), so the domain D of f is R . 2

The output of f can be made as large or small as possible; any real number r can be the output. (In fact, given any real number
r , f (0, −r) = r .) So the range R of f is R .

Example 12.1.2: Understanding a function of two variables

Let
−−−−−−−−− −
x2 y2
f (x, y) = √ 1 − − .
9 4

Find the domain and range of f .


Solution
2
2
y
The domain is all pairs (x, y) allowable as input in f . Because of the square-root, we need (x, y) such that 0 ≤ 1 − x

9

4
:
2 2
x y
0 ≤1− −
9 4
2 2
x y
+ ≤1
9 4

The above equation describes the interior of an ellipse as shown in Figure 12.1.1. We can represent the domain D graphically
2
2 y
with the figure; in set notation, we can write D = {(x, y)| x

9
+
4
≤ 1} .

Figure 12.1.1 : Illustrating the domain of f (x, y) in Example 12.1.2


The range is the set of all possible output values. The square-root ensures that all output is ≥ 0 . Since the x and y terms are
squared, then subtracted, inside the square-root, the largest output value comes at x = 0 , y = 0 : f (0, 0) = 1 . Thus the range R
is the interval [0, 1].

12.1.1 https://math.libretexts.org/@go/page/4227
Graphing Functions of Two Variables
The graph of a function f of two variables is the set of all points (x, y, f (x, y))where (x, y) is in the domain of f . This creates a
surface in space.

Figure 12.1.2 : Graphing a function of two variables.


One can begin sketching a graph by plotting points, but this has limitations. Consider Figure 12.1.2a where 25 points have been
plotted of f (x, y) = 1
. More points have been plotted than one would reasonably want to do by hand, yet it is not clear at all
x2 +y 2 +1

what the graph of the function looks like. Technology allows us to plot lots of points, connect adjacent points with lines and add
shading to create a graph like Figure 12.1.2b which does a far better job of illustrating the behavior of f .
While technology is readily available to help us graph functions of two variables, there is still a paper-and-pencil approach that is
useful to understand and master as it, combined with high-quality graphics, gives one great insight into the behavior of a function.
This technique is known as sketching level curves.

Level Curves
It may be surprising to find that the problem of representing a three dimensional surface on paper is familiar to most people (they
just don't realize it). Topographical maps, like the one shown in Figure 12.1.3, represent the surface of Earth by indicating points
with the same elevation with contour lines. The elevations marked are equally spaced; in this example, each thin line indicates an
elevation change in 50 ft increments and each thick line indicates a change of 200 ft. When lines are drawn close together,
elevation changes rapidly (as one does not have to travel far to rise 50 ft). When lines are far apart, such as near "Aspen
Campground,'' elevation changes more gradually as one has to walk farther to rise 50 ft.

Figure 12.1.3 : A topographical map displays elevation by drawing contour lines, along with the elevation is constant.

12.1.2 https://math.libretexts.org/@go/page/4227
Given a function z = f (x, y) , we can draw a "topographical map'' of f by drawing level curves (or, contour lines). A level curve at
z = c is a curve in the x-y plane such that for all points (x, y) on the curve, f (x, y) = c .

When drawing level curves, it is important that the c values are spaced equally apart as that gives the best insight to how quickly
the "elevation'' is changing. Examples will help one understand this concept.

Example 12.1.3: Drawing Level Curves


−−−−−−−−−
2 y2
Let f (x, y) = √1 − x

9

4
. Find the level curves of f for c = 0 , 0.2, 0.4, 0.6, 0.8 and 1.

Solution
−−−−−−−−−
2
y
Consider first c = 0 . The level curve for c = 0 is the set of all points (x, y) such that 0 = √1 − x2

9

4
. Squaring both sides
gives us
2 2
x y
+ = 1, (12.1.1)
9 4

an ellipse centered at (0, 0) with horizontal major axis of length 6 and minor axis of length 4. Thus for any point (x, y) on this
curve, f (x, y) = 0.
Now consider the level curve for c = 0.2
−−−−−−−−− −
2 2
x y
0.2 = √ 1 − −
9 4
2 2
x y
0.04 = 1 − −
9 4
2 2
x y
+ = 0.96
9 4
2 2
x y
+ = 1.
8.64 3.84


− −
− −
− −

This is also an ellipse, where a = √8.64 ≈ 2.94 and b = √3.84 ≈ 1.96 .
In general, for z = c , the level curve is:
−−−−−−−−− −
x2 y2
c = √1− −
9 4
2 2
2
x y
c =1− −
9 4
2 2
x y 2
+ = 1 −c
9 4
2 2
x y
+ = 1,
2 2
9(1 − c ) 4(1 − c )

ellipses that are decreasing in size as c increases. A special case is when c = 1 ; there the ellipse is just the point (0, 0).
The level curves are shown in Figure 12.4(a). Note how the level curves for c =0 and c = 0.2 are very, very close together:
this indicates that f is growing rapidly along those curves.

12.1.3 https://math.libretexts.org/@go/page/4227
Figure 12.1.4 : Graphing the level curves in Example 12.1.4
In Figure 12.1.4b, the curves are drawn on a graph of f in space. Note how the elevations are evenly spaced. Near the level
curves of c = 0 and c = 0.2 we can see that f indeed is growing quickly.

Example 12.1.4: Analyzing Level Curves


x+y
Let f (x, y) = 2
x +y
2
+1
. Find the level curves for z = c .

Solution
We begin by setting f (x, y) = c for an arbitrary c and seeing if algebraic manipulation of the equation reveals anything
significant.
x +y
=c
2 2
x +y +1
2 2
x +y = c(x +y + 1).

We recognize this as a circle, though the center and radius are not yet clear. By completing the square, we can obtain:
2 2
1 1 1
(x − ) + (y − ) = − 1, (12.1.2)
2
2c 2c 2c

−−−−−−−−− –
a circle centered at (1/(2c), 1/(2c)) with radius √1/(2c ) − 1 , where |c| < 1/√2 . The level curves for c = ±0.2, ± 0.4
2

and ±0.6 are sketched in Figure 12.1.5a. To help illustrate "elevation,'' we use thicker lines for c values near 0, and dashed
lines indicate where c < 0 .
There is one special level curve, when c = 0 . The level curve in this situation is x + y = 0 , the line y = −x .
In Figure 12.1.5b we see a graph of the surface. Note how the y -axis is pointing away from the viewer to more closely
resemble the orientation of the level curves in (a).

12.1.4 https://math.libretexts.org/@go/page/4227
Figure 12.1.5 : Graphing the level curves in Example 12.1.4
Seeing the level curves helps us understand the graph. For instance, the graph does not make it clear that one can "walk'' along
the line y = −x without elevation change, though the level curve does.

Functions of Three Variables


We extend our study of multivariable functions to functions of three variables. (One can make a function of as many variables as
one likes; we limit our study to three variables.)

Definition 78 Function of Three Variables

Let Dbe a subset of R . A function f of three variables is a rule that assigns each triple
3
(x, y, z) in D a value
w = f (x, y, z) in R . D is the domain of f ; the set of all outputs of f is the range.

Note how this definition closely resembles that of Definition 77.

Example 12.1.5: Understanding a function of three variables

Let
2
x + z + 3 sin y
f (x, y, z) = .
x + 2y − z

Evaluate f at the point (3, 0, 2) and find the domain and range of f .
Solution
2
3 + 2 + 3 sin 0
f (3, 0, 2) = = 11.
3 + 2(0) − 2

As the domain of f is not specified, we take it to be the set of all triples (x, y, z) for which f (x, y, z) is defined. As we cannot
divide by 0, we find the domain D is

D = {(x, y, z) | x + 2y − z ≠ 0}.

We recognize that the set of all points in R that are not in D form a plane in space that passes through the origin (with normal
3

vector ⟨1, 2, −1⟩).


We determine the range R is R; that is, all real numbers are possible outputs of f . There is no set way of establishing this.
Rather, to get numbers near 0 we can let y = 0 and choose z ≈ −x . To get numbers of arbitrarily large magnitude, we can let
2

z ≈ x + 2y .

12.1.5 https://math.libretexts.org/@go/page/4227
Level Surfaces
It is very difficult to produce a meaningful graph of a function of three variables. A function of one variable is a curve drawn in 2
dimensions; a function of two variables is a surface drawn in 3 dimensions; a function of three variables is a hypersurface drawn in
4 dimensions.
There are a few techniques one can employ to try to "picture'' a graph of three variables. One is an analogue of level curves: level
surfaces. Given w = f (x, y, z) , the level surface at w = c is the surface in space formed by all points (x, y, z) where
f (x, y, z) = c .

Example 12.1.6: Finding level surfaces

If a point source S is radiating energy, the intensity I at a given point P in space is inversely proportional to the square of the
distance between S and P . That is, when S = (0, 0, 0), I (x, y, z) = for some constant k .
x +y
2
k
2
+z
2

Let k = 1 ; find the level surfaces of I .

Solution
We can (mostly) answer this question using "common sense.'' If energy (say, in the form of light) is emanating from the origin,
its intensity will be the same at all points equidistant from the origin. That is, at any point on the surface of a sphere centered at
the origin, the intensity should be the same. Therefore, the level surfaces are spheres.
We now find this mathematically. The level surface at I =c is defined by
1
c = . (12.1.3)
2 2 2
x +y +z

A small amount of algebra reveals


1
2 2 2
x +y +z = . (12.1.4)
c

Given an intensity c , the level surface I =c is a sphere of radius 1/√c, centered at the origin.

Figure 12.1.6 : A table of c values and the corresponding radius r of the spheres of constant value in Example 12.1.6
Figure 12.1.6 gives a table of the radii of the spheres for given c values. Normally one would use equally spaced c values, but
these values have been chosen purposefully. At a distance of 0.25 from the point source, the intensity is 16; to move to a point
of half that intensity, one just moves out 0.1 to 0.35 - not much at all. To again halve the intensity, one moves 0.15, a little more
than before.
Note how each time the intensity if halved, the distance required to move away grows. We conclude that the closer one is to the
source, the more rapidly the intensity changes.

In the next section we apply the concepts of limits to functions of two or more variables.

This page titled 12.1: Introduction to Multivariable Functions is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated
by Gregory Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is
available upon request.

12.1.6 https://math.libretexts.org/@go/page/4227
12.2: Limits and Continuity of Multivariable Functions
We continue with the pattern we have established in this text: after defining a new kind of function, we apply calculus ideas to it.
The previous section defined functions of two and three variables; this section investigates what it means for these functions to be
"continuous.''
We begin with a series of definitions. We are used to "open intervals'' such as (1, 3), which represents the set of all x such that
1 < x < 3 , and "closed intervals'' such as [1, 3], which represents the set of all x such that 1 ≤ x ≤ 3 . We need analogous

definitions for open and closed sets in the x-y plane.

Definition 79 Open Disk, Boundary and Interior Points, Open and Closed Sets, Bounded Sets
−−−−−−−−−−−−−−−− −
An open disk B in R centered at (x
2
0, y0 ) with radius r is the set of all points (x, y) such that √(x − x
0)
2
+ (y − y0 )
2
<r .
Let S be a set of points in R . A point P in R is a boundary point of S if all open disks centered at P contain both points in
2 2

S and points not in S .

A point P in S is an interior point of S if there is an open disk centered at P that contains only points in S .
A set S is open if every point in S is an interior point.
A set S is closed if it contains all of its boundary points.
A set S is bounded if there is an M > 0 such that the open disk, centered at the origin with radius M , contains S . A set
that is not bounded is unbounded.

Figure 12.7 shows several sets in the x-y plane. In each set, point P lies on the boundary of the set as all open disks centered there
1

contain both points in, and not in, the set. In contrast, point P is an interior point for there is an open disk centered there that lies
2

entirely within the set.

Figure 12.7: Illustrating open and closed sets in the x-y plane.
The set depicted in Figure 12.7(a) is a closed set as it contains all of its boundary points. The set in (b) is open, for all of its points
are interior points (or, equivalently, it does not contain any of its boundary points). The set in (c) is neither open nor closed as it
contains some of its boundary points.

12.2.1 https://math.libretexts.org/@go/page/4228
Example 12.2.1: Determining open/closed, bounded/unbounded
−−−−−−−−−
2
y
Determine if the domain of the function f (x, y) = √1 − x2

9

4
is open, closed, or neither, and if it is bounded.

Solution
2 2
y
This domain of this function was found in Example 12.1.1 to be D = {(x, y) |
x

9
+
4
≤ 1} , the region bounded by the
2 2
y
ellipse +
x

9
= 1 . Since the region includes the boundary (indicated by the use of "≤''), the set contains all of its boundary
4

points and hence is closed. The region is bounded as a disk of radius 4, centered at the origin, contains D.

Example 12.2.2: Determining open/closed, bounded/unbounded

Determine if the domain of f (x, y) = 1

x−y
is open, closed, or neither.

Solution
As we cannot divide by 0, we find the domain to be D = {(x, y) | x − y ≠ 0} . In other words, the domain is the set of all
points (x, y) not on the line y = x .

Figure 12.8: Sketching the domain of the function in Example 12.2.2


The domain is sketched in Figure 12.8. Note how we can draw an open disk around any point in the domain that lies entirely
inside the domain, and also note how the only boundary points of the domain are the points on the line y = x . We conclude the
domain is an open set. The set is unbounded.

Limits
Recall a pseudo--definition of the limit of a function of one variable: "lim f (x) = L'' means that if x is "really close'' to c , then
x→c

f (x) is "really close'' to L. A similar pseudo--definition holds for functions of two variables. We'll say that
" lim f (x, y) = L " (12.2.1)
(x,y)→( x0 , y0 )

means "if the point (x, y) is really close to the point (x0 , y0 ) , then f (x, y) is really close to L .'' The formal definition is given
below.

Definition 80 Limit of a Function of Two Variables

Let S be an open set containing (x , y ), and let f be a function of two variables defined on
0 0 S , except possibly at (x0 , y0 ) .
The limit of f (x, y) as (x, y) approaches (x , y ) is L, denoted
0 0

lim f (x, y) = L, (12.2.2)


(x,y)→( x0 , y0 )

means that given any ϵ > 0 , there exists δ > 0 such that for all (x, y) ≠ (x0 , y0 ) , if (x, y) is in the open disk centered at
(x , y ) with radius δ , then |f (x, y) − L| < ϵ.
0 0

The concept behind Definition 80 is sketched in Figure 12.9. Given ϵ > 0 , find δ > 0 such that if (x, y) is any point in the open
disk centered at (x , y ) in the x-y plane with radius δ , then f (x, y) should be within ϵ of L.
0 0

12.2.2 https://math.libretexts.org/@go/page/4228
Figure 12.9: Illustrating the definition of a limit. The open disk in the x-y plane has radius δ. Let (x,y) be any point in this disk;
f (x, y) is within ϵ of L.

Computing limits using this definition is rather cumbersome. The following theorem allows us to evaluate limits much more easily.

THEOREM 101 Basic Limit Properties of Functions of Two Variables

Let b , x , y , L and K be real numbers, let n be a positive integer, and let f and g be functions with the following limits:
0 0

lim f (x, y) = L \ and\ lim g(x, y) = K. (12.2.3)


(x,y)→( x0 , y ) (x,y)→( x0 , y )
0 0

The following limits hold.


1. Constants: lim b =b
(x,y)→( x0 , y )
0

2. Identity : lim x = x0 ; lim y = y0


(x,y)→( x0 , y0 ) (x,y)→( x0 , y0 )

3. Sums/Differences: lim (f (x, y) ± g(x, y)) = L ± K


(x,y)→( x0 , y0 )

4. Scalar Multiples: lim b ⋅ f (x, y) = bL


(x,y)→( x0 , y0 )

5. Products: lim f (x, y) ⋅ g(x, y) = LK


(x,y)→( x0 , y )
0

6. Quotients: lim f (x, y)/g(x, y) = L/K , (K ≠ 0)


(x,y)→( x0 , y0 )

7. Powers: lim f (x, y )


n
=L
n

(x,y)→( x0 , y0 )

This theorem, combined with Theorems 2 and 3 of Section 1.3, allows us to evaluate many limits.

Example 12.2.3: Evaluating a limit

Evaluate the following limits:


y 3xy
1. lim + cos(xy) 2. lim (12.2.4)
2 2
(x,y)→(1,π) x (x,y)→(0,0) x +y

Solution
1. The aforementioned theorems allow us to simply evaluate y/x + cos(xy) when x = 1 and y = π . If an indeterminate form
is returned, we must do more work to evaluate the limit; otherwise, the result is the limit. Therefore
y π
lim + cos(xy) = + cos π
(x,y)→(1,π) x 1

= π − 1.

2. We attempt to evaluate the limit by substituting 0 in for x and y , but the result is the indeterminate form "0/0.'' To evaluate
this limit, we must "do more work,'' but we have not yet learned what "kind'' of work to do. Therefore we cannot yet
evaluate this limit.

When dealing with functions of a single variable we also considered one--sided limits and stated

12.2.3 https://math.libretexts.org/@go/page/4228
lim f (x) = L if, and only if, lim f (x) = L and lim f (x) = L. (12.2.5)
+ −
x→c x→c x→c

That is, the limit is L if and only if f (x) approaches L when x approaches c from either direction, the left or the right.
In the plane, there are infinite directions from which (x, y) might approach (x , y ). In fact, we do not have to restrict ourselves to
0 0

approaching (x , y ) from a particular direction, but rather we can approach that point along a path that is not a straight line. It is
0 0

possible to arrive at different limiting values by approaching (x , y ) along different paths. If this happens, we say that
0 0

lim f (x, y) does not exist (this is analogous to the left and right hand limits of single variable functions not being equal).
(x,y)→( x0 , y0 )

Our theorems tell us that we can evaluate most limits quite simply, without worrying about paths. When indeterminate forms arise,
the limit may or may not exist. If it does exist, it can be difficult to prove this as we need to show the same limiting value is
obtained regardless of the path chosen. The case where the limit does not exist is often easier to deal with, for we can often pick
two paths along which the limit is different.

Example 12.2.4: Showing limits do not exist


3xy
1. Show lim 2
x +y
2
does not exist by finding the limits along the lines y = mx .
(x,y)→(0,0)

sin(xy)
2. Show lim
x+y
does not exist by finding the limit along the path y = − sin x .
(x,y)→(0,0)

Solution
3xy
1. Evaluating lim
x2 +y 2
along the lines y = mx means replace all y 's with mx and evaluating the resulting limit:
(x,y)→(0,0)

2
3x(mx) 3mx
lim = lim
2 2 x→0 2 2
(x,mx)→(0,0) x + (mx ) x (m + 1)

3m
= lim
x→0 2
m +1
3m
= .
2
m +1

While the limit exists for each choice of m, we get a different limit for each choice of m. That is, along different lines we
get differing limiting values, meaning the limit does not exist.
sin(xy)
2. Let f (x, y) = x+y
. We are to show that lim f (x, y) does not exist by finding the limit along the path y = − sin x .
(x,y)→(0,0)

First, however, consider the limits found along the lines y = mx as done above.
2
sin (x(mx)) sin(m x )
lim = lim
(x,mx)→(0,0) x + mx x→0 x(m + 1)

2
sin(m x ) 1
= lim ⋅ .
x→0 x m +1

By applying L'H\^opital's Rule, we can show this limit is 0 except when m = −1 , that is, along the line y = −x . This line
is not in the domain of f , so we have found the following fact: along every line y = mx in the domain of f ,
lim f (x, y) = 0 .
(x,y)→(0,0)

Now consider the limit along the path y = − sin x :

sin ( − x sin x) sin ( − x sin x)


lim = lim (12.2.6)
(x,− sin x)→(0,0) x − sin x x→0 x − sin x

Now apply L'H\^opital's Rule twice:

12.2.4 https://math.libretexts.org/@go/page/4228
cos ( − x sin x)(− sin x − x cos x)
= lim (" = 0/0'')
x→0 1 − cos x
2
− sin ( − x sin x)(− sin x − x cos x ) + cos ( − x sin x)(−2 cos x + x sin x)
= lim
x→0 sin x

= "2/0'' ⇒ the limit does not exist.

Step back and consider what we have just discovered. Along any line y = mx in the domain of the f (x, y), the limit is 0.
However, along the path y = − sin x , which lies in the domain of f (x, y) for all x ≠ 0 , the limit does not exist. Since the
sin(xy)
limit is not the same along every path to (0, 0), we say lim
x+y
does not exist.
(x,y)→(0,0)

Example 12.2.5: Finding a limit


2 2
5x y
Let f (x, y) = 2
x +y
2
. Find lim f (x, y).
(x,y)→(0,0)

Solution
It is relatively easy to show that along any line y = mx , the limit is 0. This is not enough to prove that the limit exists, as
demonstrated in the previous example, but it tells us that if the limit does exist then it must be 0.
To prove the limit is 0, we apply Definition 80. Let ϵ>0 be given. We want to find δ >0 such that if
−−−−−−−−−−−−−− −
< δ , then |f (x, y) − 0| < ϵ .
2 2
√(x − 0 ) + (y − 0 )

−−
− 5y
2
−−−−−−
Set δ < √ϵ/5 . Note that ∣∣ 2
x +y
2


<5 for all (x, y) ≠ (0, 0), and that if √x 2
+y
2
<δ , then x
2

2
.
−−−−−−−−−−−−−− − −−−−−−
Let √(x − 0)2
+ (y − 0 )
2 2
= √x + y
2
<δ . Consider |f (x, y) − 0|:
2 2
∣ 5x y ∣
|f (x, y) − 0| = ∣ − 0∣
2 2
∣ x +y ∣

2
∣ 5y ∣
2
= ∣x ⋅ ∣
2 2
∣ x +y ∣
2
<δ ⋅5
ϵ
< ⋅5
5

= ϵ.

−−−−−−−−−−−−−− − 5x y
2 2

Thus if √(x − 0)2


+ (y − 0 )
2
<δ then |f (x, y) − 0| < ϵ , which is what we wanted to show. Thus lim 2
x +y
2
=0 .
(x,y)→(0,0)

Continuity
Definition 3 defines what it means for a function of one variable to be continuous. In brief, it meant that the graph of the function
did not have breaks, holes, jumps, etc. We define continuity for functions of two variables in a similar way as we did for functions
of one variable.

Definition 81 Continuous
Let a function f (x, y) be defined on an open disk B containing the point (x 0, y0 ) .
1. f is continuous at (x 0, y0 ) if lim f (x, y) = f (x0 , y0 ) .
(x,y)→( x0 , y0 )

2. f is continuous on B if f is continuous at all points in B . If f is continuous at all points in R , we say that f is 2

continuous everywhere.

12.2.5 https://math.libretexts.org/@go/page/4228
Example 12.2.6: Continuity of a function of two variables
cos y sin x
x ≠0
Let f (x, y) = { x
. Is f continuous at (0, 0)? Is f continuous everywhere?
cos y x =0

Solution
To determine if f is continuous at (0, 0), we need to compare lim f (x, y) to f (0, 0).
(x,y)→(0,0)

Applying the definition of f , we see that f (0, 0) = cos 0 = 1 .


We now consider the limit lim . Substituting 0 for x and y in (cos y sin x)/x returns the indeterminate form "0/0'',
f (x, y)
(x,y)→(0,0)

so we need to do more work to evaluate this limit.


Consider two related limits: lim cos y and lim
sin x

x
. The first limit does not contain x , and since cos y is
(x,y)→(0,0) (x,y)→(0,0)

continuous,

lim cos y = lim cos y = cos 0 = 1. (12.2.7)


(x,y)→(0,0) y→0

The second limit does not contain y . By Theorem 5 we can say


sin x sin x
lim = lim = 1. (12.2.8)
(x,y)→(0,0) x x→0 x

Finally, Theorem 101 of this section states that we can combine these two limits as follows:
cos y sin x sin x
lim = lim (cos y) ( )
(x,y)→(0,0) x (x,y)→(0,0) x

sin x
=( lim cos y) ( lim )
(x,y)→(0,0) (x,y)→(0,0) x

= (1)(1)

= 1.

cos y sin x
We have found that lim
x
= f (0, 0) , so f is continuous at (0, 0).
(x,y)→(0,0)

A similar analysis shows that f is continuous at all points in R . As long as x ≠ 0 , we can evaluate the limit directly; when
2

x = 0 , a similar analysis shows that the limit is cos y. Thus we can say that f is continuous everywhere. A graph of f is given

in Figure 12.10. Notice how it has no breaks, jumps, etc.

Figure 12.10: A graph of f (x, y) in Example 12.2.6.

The following theorem is very similar to Theorem 8, giving us ways to combine continuous functions to create other continuous
functions.

12.2.6 https://math.libretexts.org/@go/page/4228
THEOREM 102 Properties of Continuous Functions
Let f and g be continuous on an open disk B , let c be a real number, and let n be a positive integer. The following functions
are continuous on B .
1. Sums/Differences: f ± g
2. Constant Multiples: c ⋅ f
3. Products: f ⋅ g
4. Quotients: f /g (as longs as g ≠ 0 on B )
5. Powers: f n


6. Roots: √f (if n is even then f ≥ 0 on B ; if n is odd, then true for all values of f on B .)
n

7. Compositions: Adjust the definitions of f and g to: Let f be continuous on B , where the range of f on B is J , and let g be
a single variable function that is continuous on J . Then g ∘ f , i.e., g(f (x, y)), is continuous on B .

Example 12.2.7: Establishing continuity of a function

Let f (x, y) = sin(x 2


cos y) . Show f is continuous everywhere.

Solution
We will apply both Theorems 8 and 102. Let f (x, y) = x . Since y is not actually used in the function, and polynomials are
1
2

continuous (by Theorem 8), we conclude f is continuous everywhere. A similar statement can be made about
1

f (x, y) = cos y . Part 3 of Theorem 102 states that f = f ⋅ f


2 is continuous everywhere, and Part 7 of the theorem states the
3 1 2

composition of sine with f is continuous: that is, sin(f ) = sin(x cos y) is continuous everywhere.
3 3
2

Functions of Three Variables


The definitions and theorems given in this section can be extended in a natural way to definitions and theorems about functions of
three (or more) variables. We cover the key concepts here; some terms from Definitions 79 and 81 are not redefined but their
analogous meanings should be clear to the reader.

Definition 82 Open Balls, Limit, Continuous


1. An open ball in R centered at (x 3
0 , y0 , z0 ) with radius r is the set of all points (x, y, z) such that
−−−−−−−−−−−−−−−−−−−−−−−−− −
2 2
√(x − x0 ) + (y − y0 ) + (z − z0 )
2
=r .
2. Let D be an open set in R containing (x , y , z ) , and let f (x, y, z) be a function of three variables defined on D, except
3
0 0 0

possibly at (x , y , z ) . The limit of f (x, y, z) as (x, y, z) approaches (x , y , z ) is L, denoted


0 0 0 0 0 0

lim f (x, y, z) = L, (12.2.9)


(x,y,z)→( x0 , y0 , z0 )

means that given any ϵ > 0 , there is a δ > 0 such that for all (x, y, z) ≠ (x , y , z ) , if (x, y, z) is in the open ball centered
0 0 0

at (x , y , z ) with radius δ , then |f (x, y, z) − L| < ϵ .


0 0 0

3. Let f (x, y, z) be defined on an open ball B containing (x , y , z ) . f is continuous at (x , y , z ) if


0 0 0 0 0 0

lim f (x, y, z) = f (x , y , z ). 0 0 0
(x,y,z)→( x0 , y0 , z0 )

These definitions can also be extended naturally to apply to functions of four or more variables. Theorem 102 also applies to
function of three or more variables, allowing us to say that the function
2
x +y
−−−−−−−−−
2 2
e √y +z +3
f (x, y, z) = (12.2.10)
sin(xyz) + 5

is continuous everywhere.
When considering single variable functions, we studied limits, then continuity, then the derivative. In our current study of
multivariable functions, we have studied limits and continuity. In the next section we study derivation, which takes on a slight twist
as we are in a multivarible context.

12.2.7 https://math.libretexts.org/@go/page/4228
This page titled 12.2: Limits and Continuity of Multivariable Functions is shared under a CC BY-NC 3.0 license and was authored, remixed,
and/or curated by Gregory Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit
history is available upon request.

12.2.8 https://math.libretexts.org/@go/page/4228
12.3: Partial Derivatives
dy
Let y be a function of x. We have studied in great detail the derivative of y with respect to x, that is, , which measures the rate at
dx

which y changes with respect to x. Consider now z = f (x, y) . It makes sense to want to know how z changes with respect to x
and/or y . This section begins our investigation into these rates of change.
Consider the function z = f (x, y) = x + 2y , as graphed in Figure 12.11(a). By fixing y = 2 , we focus our attention to all points
2 2

on the surface where the y -value is 2, shown in both parts (a) and (b) of the figure. These points form a curve in space:
z = f (x, 2) = x + 8 which is a function of just one variable. We can take the derivative of z with respect to x along this curve
2

and find equations of tangent lines, etc.

Figure 12.11: By fixing y=2, the surface f (x, y) = x 2


+ 2y
2
is a curve in space.
The key notion to extract from this example is: by treating y as constant (it does not vary) we can consider how z changes with
respect to x. In a similar fashion, we can hold x constant and consider how z changes with respect to y . This is the underlying
principle of partial derivatives. We state the formal, limit--based definition first, then show how to compute these partial
derivatives without directly taking limits.

Definition 83 Partial Derivative


Let z = f (x, y) be a continuous function on an open set S in R . 2

1. The partial derivative of f with respect to x is:


f (x + h, y) − f (x, y)
fx (x, y) = lim . (12.3.1)
h→0 h

2. The partial derivative of f with respect to y is:


f (x, y + h) − f (x, y)
fy (x, y) = lim . (12.3.2)
h→0 h

Note: Alternate notations for f x (x, y) include:


∂ ∂f ∂z
f (x, y), , , and zx , (12.3.3)
∂x ∂x ∂x

with similar notations for f


y (x, y). For ease of notation, f x (x, y) is often abbreviated f . x

12.3.1 https://math.libretexts.org/@go/page/4229
Example 12.3.1: Computing partial derivatives with the limit definition

Let f (x, y) = x 2
y + 2x + y
3
. Find f x (x, y) using the limit definition.

Solution
Using Definition 83, we have:
f (x + h, y) − f (x, y)
fx (x, y) = lim
h→0 h
2 3 2 3
(x + h ) y + 2(x + h) + y − (x y + 2x + y )
= lim
h→0 h
2 2 3 2 3
(x y + 2xhy + h y + 2x + 2h + y − (x y + 2x + y )
= lim
h→0 h
2
2xhy + h y + 2h
= lim
h→0 h

= lim 2xy + hy + 2
h→0

= 2xy + 2.

We have found f x (x, y) = 2xy + 2 .

Example 12.3.1 found a partial derivative using the formal, limit--based definition. Using limits is not necessary, though, as we can
rely on our previous knowledge of derivatives to compute partial derivatives easily. When computing f (x, y), we hold y fixed -- it x

does not vary. Therefore we can compute the derivative with respect to x by treating y as a constant or coefficient.
Just as d

dx
2
(5 x ) = 10x , we compute ∂

∂x
2
(x y) = 2xy . Here we are treating y as a coefficient.

Just as d

dx
3
(5 ) = 0 , we compute ∂x
∂ 3
(y ) = 0. Here we are treating y as a constant. More examples will help make this clear.

Example 12.3.2: Finding partial derivatives

Find f x (x, y) and f y (x, y) in each of the following.


1. f (x, y) = x y + 5y − x + 7
3 2 2

2. f (x, y) = cos(x y ) + sin x 2

−−−− −
3. f (x, y) = e
2 3
x y 2
√x + 1

Solution
1. We have f (x, y) = x 3
y
2
+ 5y
2
−x +7 . Begin with f x (x, . Keep y fixed, treating it as a constant or coefficient, as
y)

appropriate:
2 2
fx (x, y) = 3 x y − 1. (12.3.4)

Note how the 5y and 7 terms go to zero. To compute f


2
y (x, y) , we hold x fixed:
3
fy (x, y) = 2 x y + 10y. (12.3.5)

Note how the −x and 7 terms go to zero.


2. We have f (x, y) = cos(x y ) + sin x . 2

Begin with f (x, y). We need to apply the Chain Rule with the cosine term; y is the coefficient of the x-term inside the
x
2

cosine function.
2 2 2 2
fx (x, y) = − sin(x y )(y ) + cos x = −y sin(x y ) + cos x. (12.3.6)

To find f (x, y), note that x is the coefficient of the y term inside of the cosine term; also note that since x is fixed, sin x
y
2

is also fixed, and we treat it as a constant.


2 2
fy (x, y) = − sin(x y )(2xy) = −2xy sin(x y ). (12.3.7)

12.3.2 https://math.libretexts.org/@go/page/4229
−−−−−
3. We have f (x, y) = e √x + 1 .
2 3
x y 2

Beginning with f (x, y), note how we need to apply the Product Rule.
x

2 3 − −−−− 2 3 1 −1/2
x y 3 2 x y 2
fx (x, y) = e (2x y )√ x + 1 + e (x + 1 ) (2x)
2
2 3
x y
2 3 −−−−− xe
3 x y 2
= 2x y e √x + 1 +
− −−−−.
√ x2 + 1

−−−−−
Note that when finding f (x, y) we do not have to apply the Product Rule; since √x
y
2
+1 does not contain y , we treat it as
fixed and hence becomes a coefficient of the e term.
2 3
x y

2 3 − −−−− 2 3 −−−−−
x y 2 2 2 2 2 x y 2
fy (x, y) = e (3 x y )√ x + 1 = 3 x y e √x + 1 . (12.3.8)

We have shown how to compute a partial derivative, but it may still not be clear what a partial derivative means. Given
z = f (x, y) , f (x, y) measures the rate at which z changes as only x varies: y is held constant.
x

Imagine standing in a rolling meadow, then beginning to walk due east. Depending on your location, you might walk up, sharply
down, or perhaps not change elevation at all. This is similar to measuring z : you are moving only east (in the "x''-direction) and
x

not north/south at all. Going back to your original location, imagine now walking due north (in the "y ''-direction). Perhaps walking
due north does not change your elevation at all. This is analogous to z = 0 : z does not change with respect to y . We can see that
y

zx and z do not have to be the same, or even similar, as it is easy to imagine circumstances where walking east means you walk
y

downhill, though walking north makes you walk uphill.


The following example helps us visualize this more.

Example 12.3.3: Evaluating partial derivatives

Let z = f (x, y) = −x 2

1

2
y
2
+ xy + 10 . Find f x (2, 1) and fy (2, 1) and interpret their meaning.

Solution
We begin by computing f x (x, y) = −2x + y and f y (x, y) = −y + x . Thus

fx (2, 1) = −3 and fy (2, 1) = 1. (12.3.9)

It is also useful to note that f (2, 1) = 7.5. What does each of these numbers mean?
Consider f (2, 1) = −3 , along with Figure 12.12(a). If one "stands'' on the surface at the point (2, 1, 7.5) and moves parallel
x

to the x-axis (i.e., only the x-value changes, not the y -value), then the instantaneous rate of change is −3. Increasing the x-
value will decrease the z -value; decreasing the x-value will increase the z -value.

12.3.3 https://math.libretexts.org/@go/page/4229
Figure 12.12: Illustrating the meaning of partial derivatives.
Now consider f (2, 1) = 1 , illustrated in Figure 12.12(b). Moving along the curve drawn on the surface, i.e., parallel to the y -
y

axis and not changing the x-values, increases the z -value instantaneously at a rate of 1. Increasing the y -value by 1 would
increase the z -value by approximately 1.
Since the magnitude of f is greater than the magnitude of f at (2, 1), it is "steeper'' in the x-direction than in the y -direction.
x y

Second Partial Derivatives


Let z = f (x, y) . We have learned to find the partial derivatives f (x, y) and f (x, y), which are each functions of x and y .
x y

Therefore we can take partial derivatives of them, each with respect to x and y . We define these "second partials'' along with the
notation, give examples, then discuss their meaning.

Definition 84 Second Partial Derivative and Mixed Partial Derivative

Let z = f (x, y) be continuous on an open set S .


1. The second partial derivative of f with respect to x then x is
2
∂ ∂f ∂ f
( ) = = ( fx ) = fxx (12.3.10)
2 x
∂x ∂x ∂x

2. The second partial derivative of f with respect to x then y is


2
∂ ∂f ∂ f
( ) = = ( fx ) = fxy (12.3.11)
y
∂y ∂x ∂y∂x

2 2
∂ f ∂ f
Similar definitions hold for ∂y
2
= fyy and ∂x∂y
= fyx .

The second partial derivatives f xy and f yx are mixed partial derivatives.

The notation of second partial derivatives gives some insight into the notation of the second derivative of a function of a single
2
d y
variable. If y = f (x), then f (x) =
′′
. The "d y'' portion means "take the derivative of y twice,'' while "dx " means "with
dx
2
2 2

respect to x both times.'' When we only know of functions of a single variable, this latter phrase seems silly: there is only one
variable to take the derivative with respect to. Now that we understand functions of multiple variables, we see the importance of
specifying which variables we are referring to.
Note: The terms in Definition 84 all depend on limits, so each definition comes with the caveat "where the limit exists.''

12.3.4 https://math.libretexts.org/@go/page/4229
Example 12.3.4: Second partial derivatives

For each of the following, find all six first and second partial derivatives. That is, find

fx , fy , fxx , fyy , fxy and fyx . (12.3.12)

1. f (x, y) = x 3
y
2
+ 2x y
3
+ cos x
3

2. f (x, y) = x

y2

3. f (x, y) = e x
sin(x y)
2

Solution
In each, we give f and f immediately and then spend time deriving the second partial derivatives.
x y

1. f (x, y) = x 3
y
2
+ 2x y
3
+ cos x
2 2 3
fx (x, y) = 3 x y + 2y − sin x
3 2
fy (x, y) = 2 x y + 6x y
∂ ∂ 2 2 3 2
fxx (x, y) = (fx ) = (3 x y + 2y − sin x) = 6x y − cos x
∂x ∂x
∂ ∂ 3 2 3
fyy (x, y) = (fy ) = (2 x y + 6x y ) = 2 x + 12xy
∂y ∂y

∂ ∂ 2 2 3 2 2
fxy (x, y) = (fx ) = (3 x y + 2y − sin x) = 6 x y + 6 y
∂y ∂y

∂ ∂ 3 2 2 2
fyx (x, y) = (fx ) = (2 x y + 6x y ) = 6 x y + 6 y
∂x ∂x
3

2. f (x, y) = x

y2
=x y
3 −2

2
3x
fx (x, y) =
y2
3
2x
fy (x, y) = − 3
y
2
∂ ∂ 3x 6x
fxx (x, y) = (fx ) = ( 2
) = 2
∂x ∂x y y
3 3
∂ ∂ 2x 6x
fyy (x, y) = (fy ) = (− 3
) = 4
∂y ∂y y y
2 2
∂ ∂ 3x 6x
fxy (x, y) = (fx ) = ( 2
) =− 3
∂y ∂y y y
3 2
∂ ∂ 2x 6x
fyx (x, y) = (fx ) = (− 3
) =− 3
∂x ∂x y y

3. f (x, y) = e sin(x y)
x 2

Because the following partial derivatives get rather long, we omit the extra notation and just give the results. In several
cases, multiple applications of the Product and Chain Rules will be necessary, followed by some basic combination of like
terms.
x 2 x 2
fx (x, y) = e sin(x y) + 2xy e cos(x y)
2 x 2
fy (x, y) = x e cos(x y)
x 2 x 2 x 2 2 2 x 2
fxx (x, y) = e sin(x y) + 4xy e cos(x y) + 2y e cos(x y) − 4 x y e sin(x y)
4 x 2
fyy (x, y) = −x e sin(x y)
2 x 2 x 2 3 x 2
fxy (x, y) = x e cos(x y) + 2x e cos(x y) − 2 x y e sin(x y)
2 x 2 x 2 3 x 2
fyx (x, y) = x e cos(x y) + 2x e cos(x y) − 2 x y e sin(x y)

Notice how in each of the three functions in Example 12.3.4, f xy = fyx . Due to the complexity of the examples, this likely is not a
coincidence. The following theorem states that it is not.

theorem 103 Mixed Partial Derivatives

Let f be defined such that f xy and f yx are continuous on an open set S . Then for each point (x, y) in S , f xy (x, y) = fyx (x, y).

Finding fxy and f yx independently and comparing the results provides a convenient way of checking our work.

Understanding Second Partial Derivatives


Now that we know how to find second partials, we investigate what they tell us.
Again we refer back to a function y = f (x) of a single variable. The second derivative of f is "the derivative of the derivative,'' or
"the rate of change of the rate of change.'' The second derivative measures how much the derivative is changing. If f (x) < 0 , then ′′

12.3.5 https://math.libretexts.org/@go/page/4229
the derivative is getting smaller (so the graph of f is concave down); if f
′′
(x) > 0 , then the derivative is growing, making the
graph of f concave up.
Now consider z = f (x, y) . Similar statements can be made about f and f as could be made about f (x) above. When taking
xx yy
′′

derivatives with respect to x twice, we measure how much f changes with respect to x. If f (x, y) < 0, it means that as x
x xx

increases, f decreases, and the graph of f will be concave down in the x-direction. Using the analogy of standing in the rolling
x

meadow used earlier in this section, f measures whether one's path is concave up/down when walking due east.
xx

Similarly, f measures the concavity in the y -direction. If f (x, y) > 0, then f is increasing with respect to y and the graph of f
yy yy y

will be concave up in the y -direction. Appealing to the rolling meadow analogy again, f measures whether one's path is concave
yy

up/down when walking due north.


We now consider the mixed partials f and f . The mixed partial f measures how much f changes with respect to y . Once
xy yx xy x

again using the rolling meadow analogy, f measures the slope if one walks due east. Looking east, begin walking north (side--
x

stepping). Is the path towards the east getting steeper? If so, f > 0 . Is the path towards the east not changing in steepness? If so,
xy

then f = 0 . A similar thing can be said about f : consider the steepness of paths heading north while side--stepping to the east.
xy yx

The following example examines these ideas with concrete numbers and graphs.

Example 12.3.5: Understanding second partial derivatives

Let z = x − y
2 2
+ xy . Evaluate the 6 first and second partial derivatives at (−1/2, 1/2) and interpret what each of these
numbers mean.

Solution
We find that:
,\quad f (x, y) = −2y + x ,\quad
fx (x, y) = 2x + y y fxx (x, y) = 2 , \quad fyy (x, y) = −2 and fxy (x, y) = fyx (x, y) = 1 .
Thus at (−1/2, 1/2) we have
fx (−1/2, 1/2) = −1/2, fy (−1/2, 1/2) = −3/2. (12.3.13)

The slope of the tangent line at (−1/2, 1/2, −1/4)in the direction of x is −1/2: if one moves from that point parallel to the
x-axis, the instantaneous rate of change will be −1/2. The slope of the tangent line at this point in the direction of y is −3/2:

if one moves from this point parallel to the y -axis, the instantaneous rate of change will be −3/2. These tangents lines are
graphed in Figure 12.13(a) and (b), respectively, where the tangent lines are drawn in a solid line.

Figure 12.13: Understanding the second partial derivatives in Example 12.3.5.

12.3.6 https://math.libretexts.org/@go/page/4229
Now consider only Figure 12.13(a). Three directed tangent lines are drawn (two are dashed), each in the direction of x; that is,
each has a slope determined by f . Note how as y increases, the slope of these lines get closer to 0. Since the slopes are all
x

negative, getting closer to 0 means the slopes are increasing. The slopes given by f are increasing as y increases, meaning x

fxy must be positive.


Since f = f , we also expect f to increase as x increases. Consider Figure 12.13(b) where again three directed tangent
xy yx y

lines are drawn, this time each in the direction of y with slopes determined by f . As x increases, the slopes become less steep y

(closer to 0). Since these are negative slopes, this means the slopes are increasing.
Thus far we have a visual understanding of f , f , and f = f . We now interpret f and f . In Figure 12.13(a), we see a
x y xy yx xx yy

curve drawn where x is held constant at x = −1/2: only y varies. This curve is clearly concave down, corresponding to the
fact that f < 0 . In part (b) of the figure, we see a similar curve where y is constant and only x varies. This curve is concave
yy

up, corresponding to the fact that f > 0 . xx

Partial Derivatives and Functions of Three Variables


The concepts underlying partial derivatives can be easily extend to more than two variables. We give some definitions and
examples in the case of three variables and trust the reader can extend these definitions to more variables if needed.

Definition 85 Partial Derivatives with Three Variables


Let w = f (x, y, z) be a continuous function on an open set S in R . 3

The partial derivative of f with respect to x is:


f (x + h, y, z) − f (x, y, z)
fx (x, y, z) = lim . (12.3.14)
h→0 h

Similar definitions hold for f y (x, y, z) and f z (x, y, z) .

By taking partial derivatives of partial derivatives, we can find second partial derivatives of f with respect to z then y , for instance,
just as before.

Example 12.3.6: Partial derivatives of functions of three variables

For each of the following, find f , f , f , f , f , and f . x y z xz yz zz

1. f (x, y, z) = x y z + x 2 3 4 2
y
2
+x z
3 3
+y z
4 4

2. f (x, y, z) = x sin(yz)
Solution
1. f x = 2x y z
3 4
+ 2x y
2
+ 3x z ;
2 3
fy = 3 x y z
2 2 4 2
+ 2x y + 4y z
3 4
;
2
fz = 4 x y z
3 3
+ 3x z
3 2
+ 4y z ;
4 3
fxz = 8x y z
3 3
+ 9x z
2 2
;
2 2 3 3 3 2 3 2 3 4 2
fyz = 12 x y z + 16 y z ; fzz = 12 x y z + 6 x z + 12 y z

2. f x = sin(yz); fy = xz cos(yz); fz = xy cos(yz) ;


2
fxz = y cos(yz); fyz = x cos(yz) − xyz sin(yz); fzz = −x y sin(xy)

Higher Order Partial Derivatives


We can continue taking partial derivatives of partial derivatives of partial derivatives of ...; we do not have to stop with second
partial derivatives. These higher order partial derivatives do not have a tidy graphical interpretation; nevertheless they are not hard
to compute and worthy of some practice. We do not formally define each higher order derivative, but rather give just a few
examples of the notation.
∂ ∂ ∂f
fxyx (x, y) = ( ( )) and (12.3.15)
∂x ∂y ∂x

∂ ∂ ∂f
fxyz (x, y, z) = ( ( )) . (12.3.16)
∂z ∂y ∂x

12.3.7 https://math.libretexts.org/@go/page/4229
Example 12.3.7: Higher order partial derivatives

a. Let f (x, y) = x y 2 2
+ sin(xy) . Find f and f
xxy yxx .
b. Let f (x, y, z) = x 3
e
xy
+ cos(z) . Find f . xyz

Solution
a. To find f xxy , we first find f , then f , then f
x xx xxy :
2 2 2
fx = 2x y + y cos(xy) fxx = 2 y −y sin(xy)

2
fxxy = 4y − 2y sin(xy) − x y cos(xy).

To find fyxx , we first find f , then f , then f


y yx yxx :
2
fy = 2 x y + x cos(xy) fyx = 4xy + cos(xy) − xy sin(xy)

2
fyxx = 4y − y sin(xy) − (y sin(xy) + x y cos(xy))

2
= 4y − 2y sin(xy) − x y cos(xy).

Note how f =f
xxy . yxx

b. To find f , we find f , then f , then f


xyz x xy xyz :
2 xy 3 xy 3 xy 3 xy 4 xy 3 xy 4 xy
fx = 3x e + x ye fxy = 3 x e +x e + x ye = 4x e + x ye

fxyz = 0.

In the previous example we saw that f =f ; this is not a coincidence. While we do not state this as a formal theorem, as long
xxy yxx

as each partial derivative is continuous, it does not matter the order in which the partial derivatives are taken. For instance,
fxxy =f xyx =f .
yxx

This can be useful at times. Had we known this, the second part of Example 12.3.7 would have been much simpler to compute.
Instead of computing f in the x, y then z orders, we could have applied the z , then x then y order (as f = f ). It is easy to
xyz xyz zxy

see that f = − sin z ; then f and f are clearly 0 as f does not contain an x or y .
z zx zxy z

A brief review of this section: partial derivatives measure the instantaneous rate of change of a multivariable function with respect
to one variable. With z = f (x, y) , the partial derivatives f and f measure the instantaneous rate of change of z when moving
x y

parallel to the x- and y -axes, respectively. How do we measure the rate of change at a point when we do not move parallel to one of
these axes? What if we move in the direction given by the vector ⟨2, 1⟩? Can we measure that rate of change? The answer is, of
course, yes, we can. This is the topic of Section 12.6. First, we need to define what it means for a function of two variables to be
differentiable.

This page titled 12.3: Partial Derivatives is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman
et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

12.3.8 https://math.libretexts.org/@go/page/4229
12.4: Differentiability and the Total Differential
We studied differentials in Section 4.4, where Definition 18 states that if y = f (x) and f is differentiable, then dy = f (x)dx . ′

One important use of this differential is in Integration by Substitution. Another important application is approximation. Let
Δx = dx represent a change in x. When dx is small, dy ≈ Δy , the change in y resulting from the change in x. Fundamental in

this understanding is this: as dx gets small, the difference between Δy and dy goes to 0. Another way of stating this: as dx goes to
0, the error in approximating Δy with dy goes to 0.
We extend this idea to functions of two variables. Let z = f (x, y) , and let Δx = dx and Δy = dy represent changes in x and y ,
respectively. Let Δz = f (x + dx, y + dy) − f (x, y) be the change in z over the change in x and y . Recalling that f and f give x y

the instantaneous rates of z -change in the x- and y -directions, respectively, we can approximate Δz with dz = f dx + f dy ; in x y

words, the total change in z is approximately the change caused by changing x plus the change caused by changing y . In a moment
we give an indication of whether or not this approximation is any good. First we give a name to dz .

Definition 86: Total Differential


Let z = f (x, y) be continuous on an open set S . Let dx and dy represent changes in x and y , respectively. Where the partial
derivatives f and f exist, the total differential of z is
x y

dz = fx (x, y)dx + fy (x, y)dy. (12.4.1)

Example 12.4.1: Finding the total differential

Let z = x 4
e
3y
. Find dz .
Solution
We compute the partial derivatives: f x = 4x e
3 3y
and f y = 3x e
4 3y
. Following Definition 86, we have
3 3y 4 3y
dz = 4 x e dx + 3 x e dy. (12.4.2)

We can approximate Δz with dz , but as with all approximations, there is error involved. A good approximation is one in which the
error is small. At a given point (x , y ), let E and E be functions of dx and dy such that E dx + E dy describes this error.
0 0 x y x y

Then
Δz = dz + Ex dx + Ey dy

= fx (x0 , y0 )dx + fy (x0 , y0 )dy + Ex dx + Ey dy.

If the approximation of Δz by dz is good, then as dx and dy get small, so does E dx + E dy . The approximation of Δz by dz is x y

even better if, as dx and dy go to 0, so do E and E . This leads us to our definition of differentiability.
x y

Definition 87: Multivariable Differentiability


Let z = f (x, y) be defined on an open set S containing (x0 , y0 ) where f (x , y ) and f (x , y ) exist. Let dz be the total
x 0 0 y 0 0

differential of z at (x , y ), let Δz = f (x
0 0 0 + dx, y0 + dy) − f (x0 , y ) , and let E
0 and E be functions of dx and dy such
x y

that
Δz = dz + Ex dx + Ey dy. (12.4.3)

1. f is differentiable at (x , y ) if, given ϵ > 0 , there is a δ > 0 such that if ||⟨dx, dy⟩|| < δ, then ||⟨E , E ⟩|| < ϵ. That is,
0 0 x y

as dx and dy go to 0, so do E and E . x y

2. f is differentiable on S if f is differentiable at every point in S . If f is differentiable on R , we say that f is 2

differentiable everywhere.

Example 12.4.2: Showing a function is differentiable

Show f (x, y) = xy + 3y is differentiable using Definition 87.


2

Solution

12.4.1 https://math.libretexts.org/@go/page/4230
We begin by finding f (x + dx, y + dy) , Δz, f and f . x y

2
f (x + dx, y + dy) = (x + dx)(y + dy) + 3(y + dy)

2 2
= xy + xdy + ydx + dxdy + 3 y + 6ydy + 3dy .

Δz = f (x + dx, y + dy) − f (x, y) , so


2
Δz = xdy + ydx + dxdy + 6ydy + 3dy . (12.4.4)

It is straightforward to compute f x =y and f


y = x + 6y . Consider once more Δz:
2
Δz = xdy + ydx + dxdy + 6ydy + 3dy (now reorder)

2
= ydx + xdy + 6ydy + dxdy + 3dy

= (y) dx + (x + 6y) dy + (dy)dx + (3dy)dy


   
fx fy Ex Ey

= fx dx + fy dy + Ex dx + Ey dy.

With E = dy and E = 3dy , it is clear that as dx and dy go to 0, E and E also go to 0. Since this did not depend on a
x y x y

specific point (x , y ), we can say that f (x, y) is differentiable for all pairs (x, y) in R , or, equivalently, that f is
0 0
2

differentiable everywhere.

Our intuitive understanding of differentiability of functions y = f (x) of one variable was that the graph of f was "smooth.'' A
similar intuitive understanding of functions z = f (x, y) of two variables is that the surface defined by f is also "smooth,'' not
containing cusps, edges, breaks, etc. The following theorem states that differentiable functions are continuous, followed by another
theorem that provides a more tangible way of determining whether a great number of functions are differentiable or not.

THEOREM 104: Continuity and Differentiability of Multivariable Functions

Let z = f (x, y) be defined on an open set S containing (x 0, y0 ) .


If f is differentiable at (x 0, y0 ) , then f is continuous at (x 0, y0 ) .

THEOREM 105: Differentiability of Multivariable Functions


Let z = f (x, y) be defined on an open set S containing (x 0, y0 ) .
If f and f are both continuous on S , then f is differentiable on S .
x y

The theorems assure us that essentially all functions that we see in the course of our studies here are differentiable (and hence
continuous) on their natural domains. There is a difference between Definition 87 and Theorem 105, though: it is possible for a
function f to be differentiable yet f and/or f is not continuous. Such strange behavior of functions is a source of delight for
x y

many mathematicians.
When f and f exist at a point but are not continuous at that point, we need to use other methods to determine whether or not f is
x y

differentiable at that point.


For instance, consider the function
xy

2 2
(x, y) ≠ (0, 0)
x +y
f (x, y) = { (12.4.5)
0 (x, y) = (0, 0)

We can find f x (0, 0) and f y (0, 0) using Definition 83:

12.4.2 https://math.libretexts.org/@go/page/4230
f (0 + h, 0) − f (0, 0)
fx (0, 0) = lim
h→0 h

0
= lim = 0;
2
h→0 h

f (0, 0 + h) − f (0, 0)
fy (0, 0) = lim
h→0 h

0
= lim = 0.
h→0 h2

Both f and f exist at (0, 0), but they are not continuous at (0, 0), as
x y

2 2 2 2
y(y −x ) x(x −y )
fx (x, y) = and fy (x, y) = (12.4.6)
2 2 2 2 2 2
(x +y ) (x +y )

are not continuous at (0, 0). (Take the limit of f as (x, y) → (0, 0) along the x- and y -axes; they give different results.) So even
x

though f and f exist at every point in the x-y plane, they are not continuous. Therefore it is possible, by Theorem 105, for f to
x y

not be differentiable.
Indeed, it is not. One can show that f is not continuous at (0, 0) (see Example 12.2.4), and by Theorem 104, this means f is not
differentiable at (0, 0).

Approximating with the Total Differential


By the definition, when f is differentiable dz is a good approximation for Δz when dx and dy are small. We give some simple
examples of how this is used here.

Example 12.4.3: Approximating with the total differential

Let z = √−
x sin y . Approximate f (4.1, 0.8).

Solution
Recognizing that π/4 ≈ 0.785 ≈ 0.8 , we can approximate f (4.1, 0.8) using f (4, π/4) . We can easily compute
– √2 –
f (4, π/4) = √4 sin(π/4) = 2 (
2
) = √2 ≈ 1.414. Without calculus, this is the best approximation we could reasonably
come up with. The total differential gives us a way of adjusting this initial approximation to hopefully get a more accurate
answer.
We let Δz = f (4.1, 0.8) − f (4, π/4). The total differential dz is approximately equal to Δz, so

f (4.1, 0.8) − f (4, π/4) ≈ dz ⇒ f (4.1, 0.8) ≈ dz + f (4, π/4). (12.4.7)

To find dz , we need f and f .


x y

sin y sin π/4


fx (x, y) = − ⇒ fx (4, π/4) = –
2 √x 2 √4

√2/2

= = √2/8.
4

− – √2
fy (x, y) = √x cos y ⇒ fy (4, π/4) = √4
2

= √2.

Approximating 4.1 with 4 gives dx = 0.1; approximating 0.8 with π/4 gives dy ≈ 0.015. Thus
dz(4, π/4) = fx (4, π/4)(0.1) + fy (4, π/4)(0.015)

√2 –
= (0.1) + √2(0.015)
8

≈ 0.039.

Returning to Equation 12.4.7, we have

12.4.3 https://math.libretexts.org/@go/page/4230
f (4.1, 0.8) ≈ 0.039 + 1.414 = 1.4531. (12.4.8)

We, of course, can compute the actual value of f (4.1, 0.8) with a calculator; the actual value, accurate to 5 places after the
decimal, is 1.45254. Obviously our approximation is quite good.

The point of the previous example was not to develop an approximation method for known functions. After all, we can very easily
compute f (4.1, 0.8)using readily available technology. Rather, it serves to illustrate how well this method of approximation works,
and to reinforce the following concept:
"New position = old position + amount of change,'' so (12.4.9)

"New position ≈ old position + approximate amount of change.'' (12.4.10)

In the previous example, we could easily compute f (4, π/4) and could approximate the amount of z -change when computing
f (4.1, 0.8), letting us approximate the new z -value.

It may be surprising to learn that it is not uncommon to know the values of f , f and f at a particular point without actually
x y

knowing the function f . The total differential gives a good method of approximating f at nearby points.

Example 12.4.4: Approximating an unknown function

Given that f (2, −3) = 6, f x (2, −3) = 1.3 and f y (2, −3) = −0.6 , approximate f (2.1, −3.03).

Solution
The total differential approximates how much f changes from the point (2, −3) to the point (2.1, −3.03). With dx = 0.1 and
dy = −0.03 , we have

dz = fx (2, −3)dx + fy (2, −3)dy

= 1.3(0.1) + (−0.6)(−0.03)

= 0.148.

The change in z is approximately 0.148, so we approximate f (2.1, −3.03) ≈ 6.148.

Error/Sensitivity Analysis
The total differential gives an approximation of the change in z given small changes in x and y . We can use this to approximate
error propagation; that is, if the input is a little off from what it should be, how far from correct will the output be? We demonstrate
this in an example.

Example 12.4.5: Sensitivity analysis

A cylindrical steel storage tank is to be built that is 10ft tall and 4ft across in diameter. It is known that the steel will
expand/contract with temperature changes; is the overall volume of the tank more sensitive to changes in the diameter or in the
height of the tank?
Solution
A cylindrical solid with height h and radius r has volume V = πr h
2
. We can view V as a function of two variables, r and h .
We can compute partial derivatives of V :
∂V ∂V
2
= Vr (r, h) = 2πrh and = Vh (r, h) = π r . (12.4.11)
∂r ∂h

The total differential is dV = (2πrh)dr + (π r )dh.


2
When h = 10 and r = 2 , we have dV = 40πdr + 4πdh .
Note that the coefficient of dr is 40π ≈ 125.7; the coefficient of dh is a tenth of that, approximately 12.57. A small change in
radius will be multiplied by 125.7, whereas a small change in height will be multiplied by 12.57. Thus the volume of the tank
is more sensitive to changes in radius than in height.

12.4.4 https://math.libretexts.org/@go/page/4230
The previous example showed that the volume of a particular tank was more sensitive to changes in radius than in height. Keep in
mind that this analysis only applies to a tank of those dimensions. A tank with a height of 1 ft and radius of 5 ft would be more
sensitive to changes in height than in radius. One could make a chart of small changes in radius and height and find exact changes
in volume given specific changes. While this provides exact numbers, it does not give as much insight as the error analysis using
the total differential.

Differentiability of Functions of Three Variables


The definition of differentiability for functions of three variables is very similar to that of functions of two variables. We again start
with the total differential.

Definition 88: Total Differential


Let w = f (x, y, z) be continuous on an open set S . Let dx, dy and dz represent changes in x, y and z , respectively. Where the
partial derivatives f , f and f exist, the total differential of w is
x y z

dz = fx (x, y, z)dx + fy (x, y, z)dy + fz (x, y, z)dz. (12.4.12)

This differential can be a good approximation of the change in w when w = f (x, y, z) is differentiable.

Definition 89: Multivariable Differentiability


Let w = f (x, y, z) be defined on an open ball B containing (x , y , z ) where f (x , y , z ), f (x , y , z
0 0 0 x 0 0 0 y 0 0 0) and
fz (x0 , y0 , z0 ) exist. Let dw be the total differential of w at (x , y , z ) , 0 0 0 let
Δw = f (x0 + dx, y + dy, z + dz) − f (x , y , z ) , and let E , E and E be functions of dx, dy and dz such that
0 0 0 0 0 x y z

Δw = dw + Ex dx + Ey dy + Ez dz (12.4.13)

1. f is differentiable at (x , y , z ) if, given ϵ > 0 , there is a δ > 0 such that if ||⟨dx, dy, dz⟩|| < δ , then
0 0 0

||⟨E , E , E ⟩|| < ϵ.


x y z

2. f is differentiable on B if f is differentiable at every point in B . If f is differentiable on R , we say that f is 3

differentiable everywhere.

Just as before, this definition gives a rigorous statement about what it means to be differentiable that is not very intuitive. We
follow it with a theorem similar to Theorem 105.

THEOREM 106: Continuity and Differentiability of Functions of Three Variables

Let w = f (x, y, z) be defined on an open ball B containing (x 0, y0 , z0 ) .


1. If f is differentiable at (x , y , z ) , then f is continuous at (x , y , z ) .
0 0 0 0 0 0

2. If f , f and f are continuous on B , then f is differentiable on B .


x y z

This set of definition and theorem extends to functions of any number of variables. The theorem again gives us a simple way of
verifying that most functions that we encounter are differentiable on their natural domains.
This section has given us a formal definition of what it means for a functions to be "differentiable,'' along with a theorem that gives
a more accessible understanding. The following sections return to notions prompted by our study of partial derivatives that make
use of the fact that most functions we encounter are differentiable.

This page titled 12.4: Differentiability and the Total Differential is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or
curated by Gregory Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history
is available upon request.

12.4.5 https://math.libretexts.org/@go/page/4230
12.5: The Multivariable Chain Rule
The Chain Rule, as learned in Section 2.5, states that d

dx
(f (g(x))) = f (g(x))g (x)
′ ′
. If t = g(x) , we can express the Chain Rule
as
df df dt
= . (12.5.1)
dx dt dx

In this section we extend the Chain Rule to functions of more than one variable.

theorem 107 Multivariable Chain Rule, Part I


Let z = f (x, y) , x = g(t) and y = h(t) , where f , g and h are differentiable functions. Then z = f (x, y) = f (g(t), h(t)) is a
function of t , and
dz df dx dy
= = fx (x, y) + fy (x, y)
dt dt dt dt

∂f dx ∂f dy
= + .
∂x dt ∂y dt

It is good to understand what the situation of z = f (x, y) , x = g(t) and y = h(t) describes. We know that z = f (x, y) describes a
surface; we also recognize that x = g(t) and y = h(t) are parametric equations for a curve in the x-y plane. Combining these
together, we are describing a curve that lies on the surface described by f . The parametric equations for this curve are x = g(t) ,
y = h(t) and z = f (g(t), h(t)) .

Consider Figure 12.14 in which a surface is drawn, along with a dashed curve in the x-y plane. Restricting f to just the points on
df
this circle gives the curve shown on the surface. The derivative dt
gives the instantaneous rate of change of f with respect to t . If
df
we consider an object traveling along this path, dt
gives the rate at which the object rises/falls.

Figure 12.14: Understanding the application of the Multivariable Chain Rule.


We now practice applying the Multivariable Chain Rule.

Example 12.5.1: Using the Multivariable Chain Rule

Let z = x2
y +x , where x = sin t and y = e . Find
5t dz

dt
using the Chain Rule.
Solution
Following Theorem 107, we find
dx dy
2 5t
fx (x, y) = 2xy + 1, fy (x, y) = x , = cos t, = 5e . (12.5.2)
dt dt

Applying the theorem, we have

12.5.1 https://math.libretexts.org/@go/page/4231
dz 2 5t
= (2xy + 1) cos t + 5 x e . (12.5.3)
dt

This may look odd, as it seems that is a function of x, y and t . Since x and y are functions of t ,
dz

dt
dz

dt
is really just a function
of t , and we can replace x with sin t and y with e : 5t

dz 2 5t 5t 5t 2
= (2xy + 1) cos t + 5 x e = (2 sin(t)e + 1) cos t + 5 e sin t. (12.5.4)
dt

The previous example can make us wonder: if we substituted for x and y at the end to show that dz

dt
is really just a function of t ,
why not substitute before differentiating, showing clearly that z is a function of t ?
That is, z = x 2
y + x = (sin t) e
2 5t
+ sin t. Applying the Chain and Product Rules, we have
dz
5t 2 5t
= 2 sin t cos t e + 5 sin te + cos t, (12.5.5)
dt

which matches the result from the example.


This may now make one wonder "What's the point? If we could already find the derivative, why learn another way of finding it?''
In some cases, applying this rule makes deriving simpler, but this is hardly the power of the Chain Rule. Rather, in the case where
z = f (x, y) , x = g(t) and y = h(t) , the Chain Rule is extremely powerful when we do not know what f , g and/or h are. It may be

hard to believe, but often in "the real world'' we know rate--of--change information (i.e., information about derivatives) without
explicitly knowing the underlying functions. The Chain Rule allows us to combine several rates of change to find another rate of
change. The Chain Rule also has theoretic use, giving us insight into the behavior of certain constructions (as we'll see in the next
section).
We demonstrate this in the next example.

Example 12.5.2: Applying the Multivarible Chain Rule

An object travels along a path on a surface. The exact path and surface are not known, but at time t = t it is known that : 0

∂z ∂z dx dy
= 5, = −2, =3 and = 7. (12.5.6)
∂x ∂y dt dt

Find dz

dt
at time t .0

Solution
The Multivariable Chain Rule states that
dz ∂z dx ∂z dy
= +
dt ∂x dt ∂y dt

= 5(3) + (−2)(7)

= 1.

By knowing certain rates--of--change information about the surface and about the path of the particle in the x-y plane, we can
determine how quickly the object is rising/falling.

We next apply the Chain Rule to solve a max/min problem.

Example 12.5.3: Applying the Multivariable Chain Rule

Consider the surface z = x + y − xy , a paraboloid, on which a particle moves with x and y coordinates given by x = cos t
2 2

and y = sin t . Find when t = 0 , and find where the particle reaches its maximum/minimum z -values.
dz

dt

12.5.2 https://math.libretexts.org/@go/page/4231
Solution
It is straightforward to compute
dx dy
fx (x, y) = 2x − y, fy (x, y) = 2y − x, = − sin t, = cos t. (12.5.7)
dt dt

Combining these according to the Chain Rule gives:


dz
= −(2x − y) sin t + (2y − x) cos t. (12.5.8)
dt

Figure 12.15: Plotting the path of a particle on a surface in Example 12.3.5


When t = 0 , x = 1 and y = 0 . Thus dz

dt
= −(2)(0) + (−1)(1) = −1 . When t = 0 , the particle is moving down, as shown in
Figure 12.15.
To find where z -value is maximized/minimized on the particle's path, we set dz

dt
=0 and solve for t :
dz
=0 = −(2x − y) sin t + (2y − x) cos t
dt

0 = −(2 cos t − sin t) sin t + (2 sin t − cos t) cos t


2 2
0 = sin t − cos t
2 2
cos t = sin t
π
t =n (for odd n)
4

We can use the First Derivative Test to find that on [0, 2π], z has reaches its absolute minimum at t = π/4 and 5π/4; it reaches
its absolute maximum at t = 3π/4 and 7π/4, as shown in Figure 12.15.

We can extend the Chain Rule to include the situation where z is a function of more than one variable, and each of these variables
is also a function of more than one variable. The basic case of this is where z = f (x, y) , and x and y are functions of two
variables, say s and t .

THEOREM 108 Multivariable Chain Rule, Part II


1. Let z = f (x, y) , x = g(s, t) and y = h(s, t) , where f , g and h are differentiable functions. Then z is a function of s and
t , and
∂f ∂f ∂y
- ∂z

∂s
=
∂x
∂x

∂s
+
∂y ∂s
, and
∂f ∂f ∂y
- ∂z

∂t
=
∂x
∂x

∂t
+
∂y ∂t
.

2. Let z = f (x , x , … , x ) be a differentiable function of m variables, where each of the x is a differentiable function of


1 2 m i

the variables t , t , … , t . Then z is a function of the t , and


1 2 n i

∂z ∂f ∂x1 ∂f ∂x2 ∂f ∂xm


= + +⋯ + . (12.5.9)
∂ti ∂x1 ∂ti ∂x2 ∂ti ∂xm ∂ti

12.5.3 https://math.libretexts.org/@go/page/4231
Example 12.5.4: Using the Multivarible Chain Rule, Part II

Let z = x2
y +x ,x=s
2
+ 3t and y = 2s − t . Find ∂z

∂s
and ∂z

∂t
, and evaluate each when s = 1 and t = 2 .

Solution
Following Theorem 108, we compute the following partial derivatives:
∂f ∂f
2
= 2xy + 1 =x , (12.5.10)
∂x ∂y

∂x ∂x ∂y ∂y
= 2s =3 =2 = −1. (12.5.11)
∂s ∂t ∂s ∂t

Thus
∂z
2 2
= (2xy + 1)(2s) + (x )(2) = 4xys + 2s + 2 x , and (12.5.12)
∂s

∂z 2 2
= (2xy + 1)(3) + (x )(−1) = 6xy − x + 3. (12.5.13)
∂t

When s = 1 and t = 2 , x = 7 and y = 0 , so


∂z ∂z
= 100 and = −46. (12.5.14)
∂s ∂t

Example 12.5.5: Using the Multivarible Chain Rule, Part II

Let w = xy + z , where x = t
2 2
e
s
, y = t cos s , and z = s sin t . Find ∂w

∂t
when s = 0 and t = π .

Solution
Following Theorem 108, we compute the following partial derivatives:
∂f ∂f ∂f
=y =x = 2z, (12.5.15)
∂x ∂y ∂z

∂x s
∂y ∂z
= 2te = cos s = s cos t. (12.5.16)
∂t ∂t ∂t

Thus
∂w s
= y(2te ) + x(cos s) + 2z(s cos t). (12.5.17)
∂t

When s = 0 and t = π , we have x = π , y = π and z = 0 . Thus


2

∂w 2 2
= π(2π) + π = 3π . (12.5.18)
∂t

12.5.4 https://math.libretexts.org/@go/page/4231
Implicit Differentiation
dy
We studied finding dx
when y is given as an implicit function of x in detail in Section 2.6. We find here that the Multivariable
dy
Chain Rule gives a simpler method of finding dx
.
dy
For instance, consider the implicit function x 2
y − xy
3
= 3. We learned to use the following steps to find dx
:

d 2 3
d
(x y − x y ) = (3)
dx dx
dy dy
2 3 2
2xy + x −y − 3x y =0
dx dx
3
dy 2xy − y
=− . (12.5.19)
dx x2 − 3x y 2

Instead of using this method, consider z = x y − x y . The implicit function above describes the level curve z = 3 . Considering x
2 3

and y as functions of x, the Multivariable Chain Rule states that


dz ∂z dx ∂z dy
= + . (12.5.20)
dx ∂x dx ∂y dx

Since z is constant (in our example, z = 3 ), dz

dx
=0 . We also know dx

dx
=1 . Equation 12.5.20 becomes

∂z ∂z dy
0 = (1) + ⇒
∂x ∂y dx

dy ∂z ∂z
=− /
dx ∂x ∂y

fx
=− .
fy

dy
Note how our solution for in Equation
dx
12.5.19 is just the partial derivative of z with respect to x , divided by the partial
derivative of z with respect to y .
We state the above as a theorem.

THEOREM 109 Implicit Differentiation


Let f be a differentiable function of x and y , where f (x, y) = c defines y as an implicit function of x, for some constant c .
Then
dy fx (x, y)
=− . (12.5.21)
dx fy (x, y)

We practice using Theorem 109 by applying it to a problem from Section 2.6.

Example 12.5.6: Implicit Differentiation

Given the implicitly defined function sin(x y ) + y = x + y , find y . Note: this is the same problem as given in Example
2 2 3 ′

2.6.4 from Section 2.6, where the solution took about a full page to find.

Solution
dy
Let f (x, y) = sin(x y ) + y − x − y ; the implicitly defined function above is equivalent to
2 2 3
f (x, y) = 0 . We find dx
by
applying Theorem 109. We find
2 2 2 2 2 2 2
fx (x, y) = 2x y cos(x y ) − 1 and fy (x, y) = 2 x y cos(x y ) + 3 y − 1, (12.5.22)

so

12.5.5 https://math.libretexts.org/@go/page/4231
2 2 2
dy 2x y cos(x y ) − 1
=− , (12.5.23)
2 2 2 2
dx 2 x y cos(x y ) + 3 y −1

which matches our solution from Example 2.6.4.

This page titled 12.5: The Multivariable Chain Rule is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by
Gregory Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available
upon request.

12.5.6 https://math.libretexts.org/@go/page/4231
12.6: Directional Derivatives
Partial derivatives give us an understanding of how a surface changes when we move in the x and y directions. We made the
comparison to standing in a rolling meadow and heading due east: the amount of rise/fall in doing so is comparable to f . x

Likewise, the rise/fall in moving due north is comparable to f . The steeper the slope, the greater in magnitude f . But what if we
y y

didn't move due north or east? What if we needed to move northeast and wanted to measure the amount of rise/fall? Partial
derivatives alone cannot measure this. This section investigates directional derivatives, which do measure this rate of change. We
begin with a definition.

Definition 90 Directional Derivatives


Let z = f (x, y) be continuous on an open set S and let u⃗ = ⟨u1 , u2 ⟩ be a unit vector. For all points (x, y), the directional
derivative of f at (x, y) in the direction of u⃗ is
f (x + h u1 , y + h u2 ) − f (x, y)
Du⃗ f (x, y) = lim . (12.6.1)
h→0 h

The partial derivatives f and f are defined with similar limits, but only x or y varies with h , not both. Here both x and y vary
x y

with a weighted h , determined by a particular unit vector u⃗. This may look a bit intimidating but in reality it is not too difficult to
deal with; it often just requires extra algebra. However, the following theorem reduces this algebraic load.

theorem 110 Directional Derivatives


Let z = f (x, y) be differentiable on an open set S containing (x 0, y0 ) , and let u⃗ = ⟨u1 , u2 ⟩ be a unit vector. The directional
derivative of f at (x , y ) in the direction of u⃗ is
0 0

D f (x0 , y0 ) = fx (x0 , y0 )u1 + fy (x0 , y0 )u2 . (12.6.2)


u⃗

Example 12.6.1: Computing directional derivatives

Let z = 14 − x 2
−y
2
and let P = (1, 2) . Find the directional derivative of f , at P , in the following directions:
1. toward the point Q = (3, 4),
2. in the direction of ⟨2, −1⟩, and
3. toward the origin.

Figure 12.16: Understanding the directional derivative in Example 12.6.1


Solution
The surface is plotted in Figure 12.16, where the point P = (1, 2) is indicated in the x, y-plane as well as the point (1, 2, 9)

which lies on the surface of f . We find that f (x, y) = −2x and f (1, 2) = −2 ; f (x, y) = −2y and f (1, 2) = −4 .
x x y y

1. Let u⃗ be the unit vector that points from the point (1, 2) to the point Q = (3, 4), as shown in the figure. The vector
1


– –
P Q = ⟨2, 2⟩ ; the unit vector in this direction is u⃗ 1 = ⟨1/ √2, 1/ √2⟩ . Thus the directional derivative of f at (1, 2) in the
direction of u⃗ is
1

– – –
Du⃗ f (1, 2) = −2(1/ √2) + (−4)(1/ √2) = −6/ √2 ≈ −4.24. (12.6.3)
1

12.6.1 https://math.libretexts.org/@go/page/4232
Thus the instantaneous rate of change in moving from the point (1, 2, 9) on the surface in the direction of u⃗ (which points 1

toward the point Q) is about −4.24. Moving in this direction moves one steeply downward.
– –
2. We seek the directional derivative in the direction of ⟨2, −1⟩. The unit vector in this direction is u⃗ = ⟨2/√5, −1/√5⟩. 2

Thus the directional derivative of f at (1, 2) in the direction of u⃗ is 2

– –
D ⃗
f (1, 2) = −2(2/ √5) + (−4)(−1/ √5) = 0. (12.6.4)
u2

Starting on the surface of f at (1, 2) and moving in the direction of ⟨2, −1⟩ (or u⃗ ) results in no instantaneous change in z -
2

value. This is analogous to standing on the side of a hill and choosing a direction to walk that does not change the elevation.
One neither walks up nor down, rather just "along the side'' of the hill.

Finding these directions of "no elevation change'' is important.


3. At P = (1, 2), the direction towards the origin is given by the vector ⟨−1, −2⟩; the unit vector in this direction is
– –
u⃗ = ⟨−1/ √5, −2/ √5⟩. The directional derivative of f at P in the direction of the origin is
3

– – –
D⃗ f (1, 2) = −2(−1/ √5) + (−4)(−2/ √5) = 10/ √5 ≈ 4.47. (12.6.5)
u3

Moving towards the origin means "walking uphill'' quite steeply, with an initial slope of about 4.47.

As we study directional derivatives, it will help to make an important connection between the unit vector u⃗ = ⟨u 1, u2 ⟩ that
describes the direction and the partial derivatives f and f . We start with a definition and follow this with a Key Idea.
x y

Definition 91 Gradient
Let z = f (x, y) be differentiable on an open set S that contains the point (x 0, y0 ) .
1. The gradient of f is ∇f (x, y) = ⟨f (x, y), f (x, y)⟩.
x y

2. The gradient of f at (x , y ) is ∇f (x , y ) = ⟨f (x , y
0 0 0 0 x 0 0 ), fy (x0 , y0 )⟩ .

Note: The symbol "∇'' is named "nabla,'' derived from the Greek name of a Jewish harp. Oddly enough, in mathematics the
expression ∇f is pronounced "del f .''
To simplify notation, we often express the gradient as ∇f = ⟨fx , fy ⟩ . The gradient allows us to compute directional derivatives in
terms of a dot product.

KEY IDEA 55 The Gradient and Directional Derivatives


The directional derivative of z = f (x, y) in the direction of u⃗ is
D ⃗ f = ∇f ⋅ u⃗. (12.6.6)
u

The properties of the dot product previously studied allow us to investigate the properties of the directional derivative. Given that
the directional derivative gives the instantaneous rate of change of z when moving in the direction of u⃗, three questions naturally
arise:
1. In what direction(s) is the change in z the greatest (i.e., the "steepest uphill'')?
2. In what direction(s) is the change in z the least (i.e., the "steepest downhill'')?
3. In what direction(s) is there no change in z ?
Using the key property of the dot product, we have
∇f ⋅ u⃗ = ∥∇f ∥ ∥u∥ cos θ = ∥∇f ∥ cos θ, (12.6.7)

where θ is the angle between the gradient and u⃗. (Since u⃗ is a unit vector, ∥u∥ = 1 .) This equation allows us to answer the three
questions stated previously.
1. Equation 12.6.7 is maximized when cos θ = 1 , i.e., when the gradient and u⃗ have the same direction. We conclude the gradient
points in the direction of greatest z change.
2. Equation 12.6.7 is minimized when cos θ = −1 , i.e., when the gradient and u⃗ have opposite directions. We conclude the
gradient points in the opposite direction of the least z change.

12.6.2 https://math.libretexts.org/@go/page/4232
3. Equation 12.6.7 is 0 when cos θ = 0 , i.e., when the gradient and u⃗ are orthogonal to each other. We conclude the gradient is
orthogonal to directions of no z change.
This result is rather amazing. Once again imagine standing in a rolling meadow and face the direction that leads you steepest uphill.
Then the direction that leads steepest downhill is directly behind you, and side--stepping either left or right (i.e., moving
perpendicularly to the direction you face) does not change your elevation at all.
Recall that a level curve is defined by a path in the xy-plane along which the z -values of a function do not change; the directional
derivative in the direction of a level curve is 0. This is analogous to walking along a path in the rolling meadow along which the
elevation does not change. The gradient at a point is orthogonal to the direction where the z does not change; i.e., the gradient is
orthogonal to level curves.
Recall that a level curve is defined as a curve in the xy-plane along which the z -values of a function do not change. Let a surface
z = f (x, y) be given, and let's represent one such level curve as a vector--valued function, r (t)
⃗ = ⟨x(t), y(t)⟩ . As the output of f

does not change along this curve, f (x(t), y(t)) = c for all t , for some constant c .
df
Since f is constant for all t , dt
=0 . By the Multivariable Chain Rule, we also know
df
′ ′
= fx (x, y)x (t) + fy (x, y)y (t)
dt
′ ′
= ⟨fx (x, y), fy (x, y)⟩ ⋅ ⟨x (t), y (t)⟩

= ∇f ⋅ r ⃗ (t)

= 0.


This last equality states ∇f ⋅ r ⃗ (t) = 0 : the gradient is orthogonal to the derivative of r ,⃗ meaning the gradient is orthogonal to r ⃗
itself. Our conclusion: at any point on a surface, the gradient at that point is orthogonal to the level curve that passes through that
point.
We restate these ideas in a theorem, then use them in an example.

theorem 111 The Gradient and Directional Derivatives


Let z = f (x, y) be differentiable on an open set S with gradient ∇f , let P = (x0 , y0 ) be a point in S and let u⃗ be a unit
vector.
1. The maximum value of D f (x , y ) is ∥∇f (x , y )∥; the direction of maximal z increase is ∇f (x , y ).
u⃗ 0 0 0 0 0 0

2. The minimum value of D f (x , y ) is −∥∇f (x , y )∥; the direction of minimal z increase is −∇f (x , y ).
u⃗ 0 0 0 0 0 0

3. At P , ∇f (x , y ) is orthogonal to the level curve passing through (x , y , f (x , y )).


0 0 0 0 0 0

Example 12.6.2: Finding directions of maximal and minimal increase

Let f (x, y) = sin x cos y and let P = (π/3, π/3). Find the directions of maximal/minimal increase, and find a direction where
the instantaneous rate of z change is 0.

Solution
We begin by finding the gradient. f x = cos x cos y and f y = − sin x sin y , thus
π π 1 3
∇f = ⟨cos x cos y, − sin x sin y⟩ and, at P , ∇f ( , ) =⟨ ,− ⟩. (12.6.8)
3 3 4 4

Thus the direction of maximal increase is ⟨1/4, −3/4⟩ . In this direction, the instantaneous rate of z change is
−−
||⟨1/4, −3/4⟩|| = √10/4 ≈ 0.79.

Figure 12.17 shows the surface plotted from two different perspectives. In each, the gradient is drawn at P with a dashed line
(because of the nature of this surface, the gradient points "into'' the surface). Let u⃗ = ⟨u , u ⟩ be the unit vector in the 1 2

direction of ∇f at P . Each graph of the figure also contains the vector ⟨u , u , ||∇f ||⟩. This vector has a "run'' of 1 (because
1 2

in the xy-plane it moves 1 unit) and a "rise'' of ||∇f ||, hence we can think of it as a vector with slope of ||∇f || in the direction
of ∇f , helping us visualize how "steep'' the surface is in its steepest direction.

12.6.3 https://math.libretexts.org/@go/page/4232
Figure 12.17: Graphing the surface and important directions in Example 12.6.2.
−−
The direction of minimal increase is ⟨−1/4, 3/4⟩; in this direction the instantaneous rate of z change is −√10/4 ≈ −0.79.
Any direction orthogonal to ∇f is a direction of no z change. We have two choices: the direction of ⟨3, 1⟩ and the direction of

⟨−3, −1⟩. The unit vector in the direction of ⟨3, 1⟩ is shown in each graph of the figure as well. The level curve at z = √3/4

is drawn: recall that along this curve the z -values do not change. Since ⟨3, 1⟩ is a direction of no z -change, this vector is
tangent to the level curve at P .

Example 12.6.3: Understanding when ∇f = 0⃗

Let f (x, y) = −x 2
+ 2x − y
2
+ 2y + 1 . Find the directional derivative of f in any direction at P = (1, 1) .
Solution
We find ∇f = ⟨−2x + 2, −2y + 2⟩ . At P , we have ∇f (1, 1) = ⟨0, 0⟩.
According to Theorem 111, this is the direction of maximal increase. However, ⟨0, 0⟩ is directionless; it has no displacement.
And regardless of the unit vector u⃗ chosen, D f = 0 . u⃗

Figure 12.18 helps us understand what this means. We can see that P lies at the top of a paraboloid. In all directions, the
instantaneous rate of change is 0.
So what is the direction of maximal increase? It is fine to give an answer of ⃗
0 = ⟨0, 0⟩ , as this indicates that all directional
derivatives are 0.

Figure 12.18: At the top of the paraboloid all directional derivatives are 0.

The fact that the gradient of a surface always points in the direction of steepest increase/decrease is very useful, as illustrated in the
following example.

12.6.4 https://math.libretexts.org/@go/page/4232
Example 12.6.4: The flow of water downhill

Consider the surface given by f (x, y) = 20 − x 2


− 2y
2
. Water is poured on the surface at (1, 1/4). What path does it take as
it flows downhill?

Solution
Let r (t)
⃗ = ⟨x(t), y(t)⟩ be the vector--valued function describing the path of the water in the xy-plane; we seek x(t) and y(t).

We know that water will always flow downhill in the steepest direction; therefore, at any point on its path, it will be moving in
the direction of −∇f . (We ignore the physical effects of momentum on the water.) Thus r ⃗ (t) will be parallel to ∇f , and there ′


is some constant c such that c∇f = r ⃗ (t) = ⟨x (t), y (t)⟩ .′ ′

dy
We find ∇f = ⟨−2x, −4y⟩ and write x (t) as ′ dx

dt
and y ′
(t) as dt
. Then
′ ′
c∇f = ⟨x (t), y (t)⟩

dx dy
⟨−2cx, −4cy⟩ = ⟨ , ⟩.
dt dt

This implies
dx dy
−2cx = and − 4cy = , i.e., (12.6.9)
dt dt

1 dx 1 dy
c =− and c =− . (12.6.10)
2x dt 4y dt

As c equals both expressions, we have


1 dx 1 dy
= . (12.6.11)
2x dt 4y dt

To find an explicit relationship between x and y , we can integrate both sides with respect to t . Recall from our study of
differentials that dt = dx . Thus:
dx

dt

1 dx 1 dy
∫ dt = ∫ dt
2x dt 4y dt

1 1
∫ dx = ∫ dy
2x 4y

1 1
ln |x| = ln |y| + C1
2 4

2 ln |x| = ln |y| + C1
2
ln | x | = ln |y| + C1

Now raise both sides as a power of e :


2 ln |y|+C1
x =e

2 ln |y| C1 C1
x =e e (Note that e is just a constant.)

2
x = yC2

1 2
x =y (Note that 1/ C2 is just a constant.)
C2
2
Cx = y.

As the water started at the point (1, 1/4), we can solve for C :

2
1 1
C (1 ) = ⇒ C = . (12.6.12)
4 4

12.6.5 https://math.libretexts.org/@go/page/4232
Figure 12.19: A graph of the surface described in Example 16.2.4 along with the path in the xy-plane with the level curves.
Thus the water follows the curve y = x /4 in the xy-plane. The surface and the path of the water is graphed in Figure
2

12.19(a). In part (b) of the figure, the level curves of the surface are plotted in the xy-plane, along with the curve y = x /4 . 2

Notice how the path intersects the level curves at right angles. As the path follows the gradient downhill, this reinforces the
fact that the gradient is orthogonal to level curves.

Functions of Three Variables


The concepts of directional derivatives and the gradient are easily extended to three (and more) variables. We combine the concepts
behind Definitions 90 and 91 and Theorem 110 into one set of definitions.

Definition 92 Directional Derivatives and Gradient with Three Variables


Let w = F (x, y, z) be differentiable on an open ball B and let u⃗ be a unit vector in R . 3

1. The gradient of F is ∇F = ⟨F , F , F ⟩ .
x y z

2. The directional derivative of F in the direction of u⃗ is


Du⃗ F = ∇F ⋅ u⃗. (12.6.13)

The same properties of the gradient given in Theorem 111, when f is a function of two variables, hold for F , a function of three
variables.

THEOREM 112 The Gradient and Directional Derivatives with Three Variables
Let w = F (x, y, z) be differentiable on an open ball B , let ∇F be the gradient of F , and let u⃗ be a unit vector.
1. The maximum value of D F is ∥∇F ∥, obtained when the angle between ∇F and u⃗ is 0, i.e., the direction of maximal
u⃗

increase is ∇F .
2. The minimum value of D F is −∥∇F ∥, obtained when the angle between ∇F and u⃗ is π, i.e., the direction of minimal
u⃗

increase is −∇F .
3. D F = 0 when ∇F and u⃗ are orthogonal.
u⃗

We interpret the third statement of the theorem as "the gradient is orthogonal to level surfaces,'' the three--variable analogue to level
curves.

Example 12.6.5: Finding directional derivatives with functions of three variables

If a point source S is radiating energy, the intensity I at a given point P in space is inversely proportional to the square of the
distance between S and P . That is, when S = (0, 0, 0), I (x, y, z) = 2
for some constant k .
x +y
k
2
+z
2

12.6.6 https://math.libretexts.org/@go/page/4232
Let k = 1 , let u⃗ = ⟨2/3, 2/3, 1/3⟩ be a unit vector, and let P = (2, 5, 3). Measure distances in inches. Find the directional
derivative of I at P in the direction of u⃗, and find the direction of greatest intensity increase at P .
Solution
We need the gradient ∇I , meaning we need I , I and I . Each partial derivative requires a simple application of the Quotient
x y z

Rule, giving
−2x −2y −2z
∇I = ⟨ , , ⟩
2 2 2 2 2 2 2 2 2 2 2 2
(x +y +z ) (x +y +z ) (x +y +z )

−4 −10 −6
∇I (2, 5, 3) = ⟨ , , ⟩ ≈ ⟨−0.003, −0.007, −0.004⟩
1444 1444 1444

Du⃗ I = ∇I (2, 5, 3) ⋅ u⃗

17
=− ≈ −0.0078.
2166

The directional derivative tells us that moving in the direction of u⃗ from P results in a decrease in intensity of about −0.008

units per inch. (The intensity is decreasing as u⃗ moves one farther from the origin than P .)
The gradient gives the direction of greatest intensity increase. Notice that
−4 −10 −6
∇I (2, 5, 3) = ⟨ , , ⟩
1444 1444 1444
2
= ⟨−2, −5, −3⟩.
1444

That is, the gradient at (2, 5, 3) is pointing in the direction of ⟨−2, −5, −3⟩, that is, towards the origin. That should make
intuitive sense: the greatest increase in intensity is found by moving towards to source of the energy.

The directional derivative allows us to find the instantaneous rate of z change in any direction at a point. We can use these
instantaneous rates of change to define lines and planes that are tangent to a surface at a point, which is the topic of the next
section.

This page titled 12.6: Directional Derivatives is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory
Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

12.6.7 https://math.libretexts.org/@go/page/4232
12.7: Tangent Lines, Normal Lines, and Tangent Planes
Derivatives and tangent lines go hand-in-hand. Given y = f (x), the line tangent to the graph of f at x = x is the line through 0

(x , f (x )) with slope f (x ) ; that is, the slope of the tangent line is the instantaneous rate of change of f at x .

0 0 0 0

When dealing with functions of two variables, the graph is no longer a curve but a surface. At a given point on the surface, it seems
there are many lines that fit our intuition of being "tangent'' to the surface.

Figure 12.20: Showing various lines tangent to a surface.


In Figures 12.20 we see lines that are tangent to curves in space. Since each curve lies on a surface, it makes sense to say that the lines
are also tangent to the surface. The next definition formally defines what it means to be "tangent to a surface.''

Definition 93 Directional Tangent Line


Let z = f (x, y) be differentiable on an open set S containing (x 0, y0 ) and let u⃗ = ⟨u 1, u2 ⟩ be a unit vector.
1. The line ℓ through (x
x 0, y0 , f (x0 , y0 )) parallel to ⟨1, 0, f (x x 0, y0 )⟩ is the tangent line to f in the direction of x at (x , y ). 0 0

2. The line ℓ through (x


y 0, y0 , f (x0 , y0 )) parallel to ⟨0, 1, f (x y 0, y0 )⟩ is the tangent line to f in the direction of y at (x , y ). 0 0

3. The line ℓ through (x


u⃗ 0, y0 , f (x0 , y0 )) parallel to ⟨u , u , D
1 2 u⃗
f (x0 , y )⟩ is the tangent line to f in the direction of u⃗ at
0

(x , y ) .
0 0

It is instructive to consider each of three directions given in the definition in terms of "slope.'' The direction of ℓ is ⟨1, 0, f (x , y )⟩; x x 0 0

that is, the "run'' is one unit in the x-direction and the "rise'' is f (x , y ) units in the z -direction. Note how the slope is just the partial
x 0 0

derivative with respect to x. A similar statement can be made for ℓ . The direction of ℓ is ⟨u , u , D f (x , y )⟩; the "run'' is one unit
y u⃗ 1 2 u⃗ 0 0

in the u⃗ direction (where u⃗ is a unit vector) and the "rise'' is the directional derivative of z in that direction.

Definition 93 leads to the following parametric equations of directional tangent lines:

⎧ x = x0 + t
⎪ ⎧ x = x0
⎪ ⎧ x = x0 + u1 t

ℓx (t) = ⎨ y = y0 , ℓy (t) = ⎨ y = y0 + t and ℓu⃗ (t) = ⎨ y = y0 + u2 t . (12.7.1)



⎪ ⎩
⎪ ⎩

z = z0 + fx (x0 , y0 )t z = z0 + fy (x0 , y0 )t z = z0 + Du⃗ f (x0 , y0 )t

Example 12.7.1: Finding directional tangent lines

Find the lines tangent to the surface z = sin x cos y at (π/2, π/2)in the x and y directions and also in the direction of v ⃗ = ⟨−1, 1⟩.
Solution
The partial derivatives with respect to x and y are:
fx (x, y) = cos x cos y ⇒ fx (π/2, π/2) = 0

fy (x, y) = − sin x sin y ⇒ fy (π/2, π/2) = −1.

At (π/2, π/2), the z -value is 0.


Thus the parametric equations of the line tangent to f at (π/2, π/2)in the directions of x and y are:

⎧ x = π/2 + t
⎪ ⎧ x = π/2

ℓx (t) = ⎨ y = π/2 and ℓy (t) = ⎨ y = π/2 + t . (12.7.2)



⎪ ⎩

z =0 z = −t

The two lines are shown with the surface in Figure 12.21(a).

12.7.1 https://math.libretexts.org/@go/page/7579
Figure 12.21: A surface and directional tangent lines in Example 12.7.1
To find the equation of the tangent line in the direction of v ⃗, we first find the unit vector in the direction of v⃗ :
– –
u⃗ = ⟨−1/ √2, 1/ √2⟩. The directional derivative at (π/2, π, 2)in the direction of u⃗ is

– – –
Du⃗ f (π/2, π, 2) = ⟨0, −1⟩ ⋅ ⟨−1/ √2, 1/ √2⟩ = −1/ √2. (12.7.3)

Thus the directional tangent line is



⎧ x = π/2 − t/ √2


ℓu⃗ (t) = ⎨ y = π/2 + t/ √2 . (12.7.4)

⎪ –
z = −t/ √2

The curve through (π/2, π/2, 0)in the direction of v ⃗ is shown in Figure 12.21(b) along with ℓ u⃗
(t) .

Example 12.7.2: Finding directional tangent lines

Let f (x, y) = 4xy − x 4


−y
4
. Find the equations of all directional tangent lines to f at (1, 1).
Solution
First note that f (1, 1) = 2 . We need to compute directional derivatives, so we need ∇f . We begin by computing partial derivatives.
3 3
fx = 4y − 4 x ⇒ fx (1, 1) = 0; fy = 4x − 4 y ⇒ fy (1, 1) = 0. (12.7.5)

Thus ∇f (1, 1) = ⟨0, 0⟩ . Let u⃗ = ⟨u1 , u2 ⟩ be any unit vector. The directional derivative of f at (1, 1) will be
D
u⃗
f (1, 1) = ⟨0, 0⟩ ⋅ ⟨u1 , u2 ⟩ = 0 . It does not matter what direction we choose; the directional derivative is always 0. Therefore

⎧ x = 1 + u1 t

ℓu⃗ (t) = ⎨ y = 1 + u2 t (12.7.6)




z =2

Figure 12.22 shows a graph of f and the point (1, 1, 2). Note that this point comes at the top of a "hill,'' and therefore every tangent
line through this point will have a "slope'' of 0.

12.7.2 https://math.libretexts.org/@go/page/7579
Figure 12.22: Graphing f in Example 12.7.2
That is, consider any curve on the surface that goes through this point. Each curve will have a relative maximum at this point, hence
its tangent line will have a slope of 0. The following section investigates the points on surfaces where all tangent lines have a slope
of 0.

Normal Lines
When dealing with a function y = f (x) of one variable, we stated that a line through (c, f (c)) was tangent to f if the line had a slope
of f (c) and was normal (or, perpendicular, orthogonal) to f if it had a slope of −1/f (c). We extend the concept of normal, or
′ ′

orthogonal, to functions of two variables.


Let be a differentiable function of two variables. By Definition 93, at (x , y ), ℓ (t) is a line parallel to the vector
z = f (x, y) 0 0 x

= ⟨1, 0, f (x , y )⟩ and ℓ (t) is a line parallel to d = ⟨0, 1, f (x , y )⟩ . Since lines in these directions through (x , y , f (x , y ))
⃗ ⃗
dx x 0 0 y y y 0 0 0 0 0 0

are tangent to the surface, a line through this point and orthogonal to these directions would be orthogonal, or normal, to the surface.
We can use this direction to create a normal line.

The direction of the normal line is orthogonal to d



x and d

y , hence the direction is parallel to d

n =d

x ×d

y . It turns out this cross
product has a very simple form:
⃗ ⃗
d x ×d y = ⟨1, 0, fx ⟩ × ⟨0, 1, fy ⟩ = ⟨−fx , −fy , 1⟩. (12.7.7)

It is often more convenient to refer to the opposite of this direction, namely ⟨f x, fy , −1⟩ . This leads to a definition.

Definition 94 Normal Line


Let z = f (x, y) be differentiable on an open set S containing (x 0, y0 ) where
a = fx (x0 , y0 ) and b = fy (x0 , y0 ) (12.7.8)

are defined.
1. A nonzero vector parallel to n⃗ = ⟨a, b, −1⟩ is orthogonal to f at P = (x , y , f (x 0 0 0, y0 )) .
2. The line ℓ through P with direction parallel to n⃗ is the normal line to f at P .
n

Thus the parametric equations of the normal line to a surface f at (x 0, y0 , f (x0 , y0 )) is:

⎧ x = x0 + at

ℓn (t) = ⎨ y = y0 + bt . (12.7.9)


z = f (x0 , y0 ) − t

Example 12.7.3: Finding a normal line

Find the equation of the normal line to z = −x 2


−y
2
+2 at (0, 1).

Solution
We find z (x, y) = −2x and z (x, y) = −2y ; at (0, 1), we have z = 0 and z = −2 . We take the direction of the normal line,
x y x y

following Definition 94, to be n⃗ = ⟨0, −2, −1⟩. The line with this direction going through the point (0, 1, 1) is

12.7.3 https://math.libretexts.org/@go/page/7579
⎧ x =0

ℓn (t) = ⎨ y = −2t + 1 or ℓn (t) = ⟨0, −2, −1⟩t + ⟨0, 1, 1⟩. (12.7.10)




z = −t + 1

Figure 12.23: Graphing a surface with a normal line from Example 12.7.3
The surface z = −x 2
+y
2
, along with the found normal line, is graphed in Figure 12.23.

The direction of the normal line has many uses, one of which is the definition of the tangent plane which we define shortly. Another
use is in measuring distances from the surface to a point. Given a point Q in space, it is general geometric concept to define the distance
¯
¯¯¯¯¯¯
¯
from Q to the surface as being the length of the shortest line segment P Q over all points P on the surface. This, in turn, implies that

PQ will be orthogonal to the surface at P . Therefore we can measure the distance from Q to the surface f by finding a point P on the

surface such that P Q is parallel to the normal line to f at P .

Example 12.7.4: Finding the distance from a point to a surface

Let f (x, y) = 2 − x 2
−y
2
and let Q = (2, 2, 2). Find the distance from Q to the surface defined by f .

Solution

This surface is used in Example 12.7.2, so we know that at (x, y), the direction of the normal line will be d ⃗ n = ⟨−2x, −2y, −1⟩ .A
→ →
point P on the surface will have coordinates (x, y, 2 − x 2
−y )
2
, so P Q = ⟨2 − x, 2 − y, x 2 2
+y ⟩ . To find where P Q is parallel

to d ⃗ , we need to find x, y and c such that cP Q = d ⃗ .
n n



cP Q = d n

2 2
c⟨2 − x, 2 − y, x +y ⟩ = ⟨−2x, −2y, −1⟩.

This implies

c(2 − x) = −2x

c(2 − y) = −2y

2 2
c(x + y ) = −1

In each equation, we can solve for c :


−2x −2y −1
c = = = . (12.7.11)
2 2
2 −x 2 −y x +y

The first two fractions imply x = y , and so the last fraction can be rewritten as c = −1/(2x ). Then 2

−2x −1
=
2
2 −x 2x
2
−2x(2 x ) = −1(2 − x)

3
4x = 2 −x
3
4x + x − 2 = 0.

This last equation is a cubic, which is not difficult to solve with a numeric solver. We find that x = 0.689 , hence
. We find the distance from Q to the surface of f is
P = (0.689, 0.689, 1.051)

12.7.4 https://math.libretexts.org/@go/page/7579
→ −−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−
2 2 2
∥ P Q∥ = √ (2 − 0.689 ) + (2 − 0.689 ) + (2 − 1.051 ) = 2.083. (12.7.12)

We can take the concept of measuring the distance from a point to a surface to find a point Q a particular distance from a surface at a
given point P on the surface.

Example 12.7.5: Finding a point a set distance from a surface

Let f (x, y) = x − y 2
+3 . Let P = (2, 1, f (2, 1)) = (2, 1, 4) . Find points Q in space that are 4 units from the surface of f at P .
→ →
That is, find Q such that ∥P Q∥ = 4 and P Q is orthogonal to f at P .
Solution
We begin by finding partial derivatives:
fx (x, y) = 1 ⇒ fx (2, 1) = 1

fy (x, y) = −2y ⇒ fy (2, 1) = −2

The vector n⃗ = ⟨1, −2, −1⟩ is orthogonal to f at P . For reasons that will become more clear in a moment, we find the unit vector
in the direction of n⃗ :
n⃗ – – –
u⃗ = = ⟨1/ √6, −2/ √6, −1/ √6⟩ ≈ ⟨0.408, −0.816, −0.408⟩. (12.7.13)
∥n∥

Thus a the normal line to f at P can be written as

ℓn (t) = ⟨2, 1, 4⟩ + t⟨0.408, −0.816, −0.408⟩. (12.7.14)

An advantage of this parametrization of the line is that letting t = t gives a point on the line that is |t | units from
0 0 P . (This is
because the direction of the line is given in terms of a unit vector.) There are thus two points in space 4 units from P :

Q1 = ℓn (4) Q2 = ℓn (−4)

≈ ⟨3.63, −2.27, 2.37⟩ ≈ ⟨0.37, 4.27, 5.63⟩

Figure 12.24: Graphing the surface in Example 12.7.5 along with points 4 units from the surface.
The surface is graphed along with points P , Q , Q and a portion of the normal line to f at P .
1 2

Tangent Planes
We can use the direction of the normal line to define a plane. With a = f (x , y ) , b = f (x , y ) and P = (x , y , f (x
x 0 0 y 0 0 0 0 0, y0 )) , the
vector n⃗ = ⟨a, b, −1⟩ is orthogonal to f at P . The plane through P with normal vector n⃗ is therefore tangent to f at P .

Definition 95 Tangent Plane


Let z = f (x, y) be differentiable on an open set S containing (x 0, y0 ) , where a = f x (x0 , y0 ) ,b=f y (x0 , y0 ) , n⃗ = ⟨a, b, −1⟩ and
P = (x , y , f (x , y )) .
0 0 0 0

The plane through P with normal vector n⃗ is the tangent plane to f at P . The standard form of this plane is

a(x − x0 ) + b(y − y0 ) − (z − f (x0 , y0 )) = 0. (12.7.15)

12.7.5 https://math.libretexts.org/@go/page/7579
Example 12.7.6: Finding tangent planes

Find the equation of the tangent plane to z = −x 2


−y
2
+2 at (0, 1).
Solution
Note that this is the same surface and point used in Example 12.7.3. There we found n⃗ = ⟨0, −2, −1⟩ and P = (0, 1, 1) . Therefore
the equation of the tangent plane is

−2(y − 1) − (z − 1) = 0. (12.7.16)

Figure 12.25: Graphing a surface with tangent plane from Example 17.2.6
The surface z = −x 2
+y
2
and tangent plane are graphed in Figure 12.25.

Example 12.7.7: Using the tangent plane to approximate function values

The point (3, −1, 4) lies on the surface of an unknown differentiable function f where f (3, −1) = 2 and f x y (3, −1) = −1/2 . Find
the equation of the tangent plane to f at P , and use this to approximate the value of f (2.9, −0.8).
Solution
Knowing the partial derivatives at (3, −1) allows us to form the normal vector to the tangent plane, n⃗ = ⟨2, −1/2, −1⟩ . Thus the
equation of the tangent line to f at P is:
2(x − 3) − 1/2(y + 1) − (z − 4) = 0 ⇒ z = 2(x − 3) − 1/2(y + 1) + 4. (12.7.17)

Just as tangent lines provide excellent approximations of curves near their point of intersection, tangent planes provide excellent
approximations of surfaces near their point of intersection. So f (2.9, −0.8) ≈ z(2.9, −0.8) = 3.7.
This is not a new method of approximation. Compare the right hand expression for z in Equation 12.7.17 to the total differential:
dz = fx dx + fy dy and z = 2 (x − 3) + −1/2 (y + 1) + 4. (12.7.18)
   
fx
dx fy dy


dz

Thus the "new z -value'' is the sum of the change in z (i.e., dz ) and the old z -value (4). As mentioned when studying the total
differential, it is not uncommon to know partial derivative information about a unknown function, and tangent planes are used to
give accurate approximations of the function.

The Gradient and Normal Lines, Tangent Planes


The methods developed in this section so far give a straightforward method of finding equations of normal lines and tangent planes for
surfaces with explicit equations of the form z = f (x, y) . However, they do not handle implicit equations well, such as
x + y + z = 1 . There is a technique that allows us to find vectors orthogonal to these surfaces based on the gradient.
2 2 2

Definition 96 Gradient
Let w = F (x, y, z) be differentiable on an open ball B that contains the point (x 0, y0 , z0 ) .
1. The gradient of F is ∇F (x, y, z) = ⟨f x (x, y, z), fy (x, y, z), fz (x, y, z)⟩ .
2. The gradient of F at (x , y , z ) is
0 0 0

∇F (x0 , y0 , z0 ) = ⟨fx (x0 , y0 , z0 ), fy (x0 , y0 , z0 ), fz (x0 , y0 , z0 )⟩. (12.7.19)

12.7.6 https://math.libretexts.org/@go/page/7579
Recall that when z = f (x, y) , the gradient ∇f = ⟨f , f ⟩ is orthogonal to level curves of f . An analogous statement can be made
x y

about the gradient ∇F , where w = F (x, y, z). Given a point (x , y , z ) , let c = F (x , y , z ) . Then F (x, y, z) = c is a level surface
0 0 0 0 0 0

that contains the point (x , y , z ) . The following theorem states that ∇F (x , y , z ) is orthogonal to this level surface.
0 0 0 0 0 0

THEOREM 113 The Gradient and Level Surfaces

Let w = F (x, y, z) be differentiable on an open ball B containing (x 0, y0 , z0 ) with gradient ∇F , where F (x 0, y0 , z0 ) = c .


The vector ∇F (x 0, y0 , z0 ) is orthogonal to the level surface F (x, y, z) = c at (x 0, y0 , z0 ) .

The gradient at a point gives a vector orthogonal to the surface at that point. This direction can be used to find tangent planes and
normal lines.

Example 12.7.8: Using the gradient to find a tangent plane


2
2 2
y
Find the equation of the plane tangent to the ellipsoid x

12
+
6
+
z

4
=1 at P = (1, 2, 1) .

Solution
2
2 y 2

We consider the equation of the ellipsoid as a level surface of a function F of three variables, where F (x, y, z) = x

12
+
6
+
z

4
.
The gradient is:
∇F (x, y, z) = ⟨Fx , Fy , Fz ⟩

x y z
=⟨ , , ⟩.
6 3 2

At P , the gradient is ∇F (1, 2, 1) = ⟨1/6, 2/3, 1/2⟩. Thus the equation of the plane tangent to the ellipsoid at P is
1 2 1
(x − 1) + (y − 2) + (z − 1) = 0. (12.7.20)
6 3 2

Figure 12.26: An ellipsoid and its tangent plane at a point.


The ellipsoid and tangent plane are graphed in Figure 12.26.

Tangent lines and planes to surfaces have many uses, including the study of instantaneous rates of changes and making approximations.
Normal lines also have many uses. In this section we focused on using them to measure distances from a surface. Another interesting
application is in computer graphics, where the effects of light on a surface are determined using normal vectors.
The next section investigates another use of partial derivatives: determining relative extrema. When dealing with functions of the form
y = f (x), we found relative extrema by finding x where f (x) = 0 . We can start finding relative extrema of z = f (x, y) by setting f

x

and f to 0, but it turns out that there is more to consider.


y

12.7: Tangent Lines, Normal Lines, and Tangent Planes is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by
LibreTexts.

12.7.7 https://math.libretexts.org/@go/page/7579
12.8: Extreme Values
Given a function z = f (x, y) , we are often interested in points where z takes on the largest or smallest values. For instance, if z
represents a cost function, we would likely want to know what (x, y) values minimize the cost. If z represents the ratio of a volume
to surface area, we would likely want to know where z is greatest. This leads to the following definition.

Definition 97 Relative and Absolute Extrema


Let z = f (x, y) be defined on a set S containing the point P = (x0 , y0 ) .
1. If there is an open disk D containing P such that f (x , y ) ≥ f (x, y) for all (x, y) in D, then f has a relative maximum
0 0

at P ; if f (x , y ) ≤ f (x, y) for all (x, y) in D, then f has a relative minimum at P .


0 0

2. If f (x , y ) ≥ f (x, y) for all (x, y) in S , then f has an absolute maximum at P ; if f (x , y ) ≤ f (x, y) for all (x, y) in S ,
0 0 0 0

then f has an absolute minimum at P .


3. If f has a relative maximum or minimum at P , then f has a relative extrema at P ; if f has an absolute maximum or
minimum at P , then f has an absolute extrema at P .

If f has a relative or absolute maximum at P = (x , y ) , it means every curve on the surface of f through P will also have a
0 0

relative or absolute maximum at P . Recalling what we learned in Section 3.1, the slopes of the tangent lines to these curves at P
must be 0 or undefined. Since directional derivatives are computed using f and f , we are led to the following definition and
x y

theorem.

Definition 98 Critical Point


Let z = f (x, y) be continuous on an open set S . A critical point P = (x0 , y0 ) of f is a point in S such that
fx (x0 , y0 ) = 0 and f (x , y ) = 0 , or
y 0 0

fx (x0 , y0 ) and/or f (x , y ) is undefined.


y 0 0

theorem 114 Critical Points and Relative Extrema


Let z = f (x, y) be defined on an open set S containing P = (x0 , y0 ) . If f has a relative extrema at P , then P is a critical
point of f .

Therefore, to find relative extrema, we find the critical points of f and determine which correspond to relative maxima, relative
minima, or neither. The following examples demonstrate this process.

Example 12.8.1: Finding critical points and relative extrema

Let f (x, y) = x 2
+y
2
− xy − x − 2 . Find the relative extrema of f .

Solution
We start by computing the partial derivatives of f :
fx (x, y) = 2x − y − 1 and fy (x, y) = 2y − x. (12.8.1)

Each is never undefined. A critical point occurs when f and f are simultaneously 0, leading us to solve the following system
x y

of linear equations:
2x − y − 1 = 0 and − x + 2y = 0. (12.8.2)

This solution to this system is x = 2/3, y = 1/3. (Check that at (2/3, 1/3), both f and f are 0.) x y

12.8.1 https://math.libretexts.org/@go/page/7580
Figure 12.27: The surface in Example 12.8.1 with its absolute minimum indicated.
The graph in Figure 12.27 shows f along with this critical point. It is clear from the graph that this is a relative minimum;
further consideration of the function shows that this is actually the absolute minimum.

Example 12.8.2: Finding critical points and relative extrema


−−−−−−
Let f (x, y) = −√x 2
+y
2
+2 . Find the relative extrema of f .

Solution
We start by computing the partial derivatives of f :
−x −y
fx (x, y) = and fy (x, y) = . (12.8.3)
−−−−−− −−−−−−
2 2 2 2
√x + y √x + y

It is clear that f = 0 when x = 0 \& y ≠ 0 , and that f = 0 when y = 0 \& x ≠ 0 . At (0, 0), both f and f are not 0, but
x y x y

rather undefined. The point (0, 0) is still a critical point, though, because the partial derivatives are undefined. This is the only
critical point of f .

Figure 12.28: The surface in Example 12.8.2 with its absolute maximum indicated.
The surface of f is graphed in Figure 12.28 along with the point . The graph shows that this point is the absolute
(0, 0, 2)

maximum of f .

In each of the previous two examples, we found a critical point of f and then determined whether or not it was a relative (or
absolute) maximum or minimum by graphing. It would be nice to be able to determine whether a critical point corresponded to a
max or a min without a graph. Before we develop such a test, we do one more example that sheds more light on the issues our test
needs to consider.

Example 12.8.3: Finding critical points and relative extrema

Let f (x, y) = x3
− 3x − y
2
+ 4y . Find the relative extrema of f .

Solution
Once again we start by finding the partial derivatives of f :

12.8.2 https://math.libretexts.org/@go/page/7580
2
fx (x, y) = 3 x −3 and fy (x, y) = −2y + 4. (12.8.4)

Each is always defined. Setting each equal to 0 and solving for x and y , we find
fx (x, y) = 0 ⇒ x = ±1

fy (x, y) = 0 ⇒ y = 2.

We have two critical points: (−1, 2) and (1, 2) . To determine if they correspond to a relative maximum or minimum, we
consider the graph of f in Figure 12.29.

Figure 12.29: The surface in Example 12.8.3 with both critical points marked.
The critical point (−1, 2) clearly corresponds to a relative maximum. However, the critical point at (1, 2) is neither a
maximum nor a minimum, displaying a different, interesting characteristic.
If one walks parallel to the y -axis towards this critical point, then this point becomes a relative maximum along this path. But if
one walks towards this point parallel to the x-axis, this point becomes a relative minimum along this path. A point that seems
to act as both a max and a min is a saddle point. A formal definition follows.

Definition 99 Saddle Point


Let P = (x , y ) be in the domain of f where f
0 0 x =0 and f = 0 at P . P is a saddle point of f if, for every open disk
y D

containing P , there are points (x , y ) and (x , y


1 1 2 2) in D such that f (x , y ) > f (x , y ) and f (x , y ) < f (x , y ) .
0 0 1 1 0 0 2 2

At a saddle point, the instantaneous rate of change in all directions is 0 and there are points nearby with z -values both less than and
greater than the z -value of the saddle point.
Before Example 12.8.3 we mentioned the need for a test to differentiate between relative maxima and minima. We now recognize
that our test also needs to account for saddle points. To do so, we consider the second partial derivatives of f .
Recall that with single variable functions, such as y = f (x), if f (c) > 0 , then f is concave up at c , and if f (c) = 0 , then f has a
′′ ′

relative minimum at x = c . (We called this the Second Derivative Test.) Note that at a saddle point, it seems the graph is "both''
concave up and concave down, depending on which direction you are considering.
It would be nice if the following were true:

However, this is not the case. Functions f exist where f and f are both positive but a saddle point still exists. In such a case,
xx yy

while the concavity in the x-direction is up (i.e., f > 0 ) and the concavity in the y -direction is also up (i.e., f > 0 ), the
xx yy

concavity switches somewhere in between the x- and y -directions.


To account for this, consider D = f f − f f . Since f and f are equal when continuous (refer back to Theorem 103), we
xx yy xy yx xy yx

can rewrite this as D = f f − f . D can be used to test whether the concavity at a point changes depending on direction. If
xx yy
2
xy

D > 0 , the concavity does not switch (i.e., at that point, the graph is concave up or down in all directions). If D < 0 , the concavity

does switch. If D = 0 , our test fails to determine whether concavity switches or not. We state the use of D in the following
theorem.

12.8.3 https://math.libretexts.org/@go/page/7580
THEOREM 115 Second Derivative Test
Let z = f (x, y) be differentiable on an open set containing P = (x0 , y0 ) , and let
2
D = fxx (x0 , y0 )fyy (x0 , y0 ) − fxy (x0 , y0 ). (12.8.5)

1. If D > 0 and f (x , y ) > 0 , then P is a relative minimum of f .


xx 0 0

2. If D > 0 and f (x , y ) < 0 , then P is a relative maximum of f .


xx 0 0

3. If D < 0 , then P is a saddle point of f .


4. If D = 0 , the test is inconclusive.

We first practice using this test with the function in the previous example, where we visually determined we had a relative
maximum and a saddle point.

Example 12.8.4: Using the Second Derivative Test

Let f (x, y) = x − 3x − y + 4y as in Example 12.8.3. Determine whether the function has a relative minimum, maximum,
3 2

or saddle point at each critical point.

Solution
We determined previously that the critical points of f are (−1, 2) and (1, 2). To use the Second Derivative Test, we must find
the second partial derivatives of f :
fxx = 6x; fyy = −2; fxy = 0. (12.8.6)

Thus D(x, y) = −12x.


At (−1, 2): D(−1, 2) = 12 > 0 , and f xx (−1, 2) = −6 . By the Second Derivative Test, f has a relative maximum at (−1, 2).
At (1, 2): D(1, 2) = −12 < 0 . The Second Derivative Test states that f has a saddle point at (1, 2).
The Second Derivative Test confirmed what we determined visually.

Example 12.8.5: Using the Second Derivative Test

Find the relative extrema of f (x, y) = x 2


y +y
2
+ xy .

Solution
We start by finding the first and second partial derivatives of f :
2
fx = 2xy + y fy = x + 2y + x

fxx = 2y fyy = 2 (12.8.7)

fxy = 2x + 1 fyx = 2x + 1.

We find the critical points by finding where fx and f are simultaneously 0 (they are both never undefined). Setting
y fx = 0 ,
we have:
fx = 0 ⇒ 2xy + y = 0 ⇒ y(2x + 1) = 0. (12.8.8)

This implies that for f x =0 , either y = 0 or 2x + 1 = 0 .


Assume y = 0 then consider f y =0 :

fy = 0

2
x + 2y + x = 0, and since y = 0, we have
2
x +x = 0

x(x + 1) = 0.

Thus if y = 0 , we have either x = 0 or x = −1 , giving two critical points: (−1, 0) and (0, 0).

12.8.4 https://math.libretexts.org/@go/page/7580
Going back to f , now assume 2x + 1 = 0 , i.e., that x = −1/2, then consider f
x y =0 :
fy = 0

2
x + 2y + x = 0, and since x = −1/2, we have

1/4 + 2y − 1/2 = 0

y = 1/8.

Thus if x = −1/2, y = 1/8 giving the critical point (−1/2, 1/8).


With D = 4y − (2x + 1) , we apply the Second Derivative Test to each critical point.
2

At (−1, 0), D < 0 , so (−1, 0) is a saddle point.


At (0, 0), D < 0 , so (0, 0) is also a saddle point.
At (−1/2, 1/8), D > 0 and f xx >0 , so (−1/2, 1/8) is a relative minimum.

Figure 12.30: Graphing f from Example 12.8.5 and its relative extrema.
Figure 12.30 shows a graph of f and the three critical points. Note how this function does not vary much near the critical
points -- that is, visually it is difficult to determine whether a point is a saddle point or relative minimum (or even a critical
point at all!). This is one reason why the Second Derivative Test is so important to have.

Constrained Optimization
When optimizing functions of one variable such as y = f (x), we made use of Theorem 25, the Extreme Value Theorem, that said
that over a closed interval I , a continuous function has both a maximum and minimum value. To find these maximum and
minimum values, we evaluated f at all critical points in the interval, as well as at the endpoints (the "boundary'') of the interval.
A similar theorem and procedure applies to functions of two variables. A continuous function over a closed set also attains a
maximum and minimum value (see the following theorem). We can find these values by evaluating the function at the critical
values in the set and over the boundary of the set. After formally stating this extreme value theorem, we give examples.

theorem 116 Extreme Value Theorem

Let z = f (x, y) be a continuous function on a closed, bounded set S . Then f has a maximum and minimum value on S .

Example 12.8.6: Finding extrema on a closed set

Let f (x, y) = x − y + 5 and let


2 2
S be the triangle with vertices ,
(−1, −2) (0, 1) and (2, −2). Find the maximum and
minimum values of f on S .

Solution
It can help to see a graph of f along with the set S . In Figure 12.31(a) the triangle defining S is shown in the xy-plane in a
dashed line. Above it is the surface of f ; we are only concerned with the portion of f enclosed by the "triangle'' on its surface.

12.8.5 https://math.libretexts.org/@go/page/7580
Figure 12.31: Plotting the surface of f along with the restricted domain S.
We begin by finding the critical points of f . With f x = 2x and fy = −2y , we find only one critical point, at (0, 0).
We now find the maximum and minimum values that f attains along the boundary of S , that is, along the edges of the triangle.
In Figure 12.31(b) we see the triangle sketched in the plane with the equations of the lines forming its edges labeled.
Start with the bottom edge, along the line y = −2 . If y is −2, then on the surface, we are considering points f (x, −2); that is,
our function reduces to f (x, −2) = x − (−2) + 5 = x + 1 = f (x) . We want to maximize/minimize f (x) = x + 1 on
2 2 2
1 1
2

the interval [−1, 2]. To do so, we evaluate f (x) at its critical points and at the endpoints.
1

The critical points of f are found by setting its derivative equal to 0:


1


f (x) = 0 ⇒ x = 0. (12.8.9)
1

Evaluating f at this critical point, and at the endpoints of [−1, 2] gives:


1

f1 (−1) = 2 ⇒ f (−1, −2) = 2

f1 (0) = 1 ⇒ f (0, −2) = 1

f1 (2) = 5 ⇒ f (2, −2) = 5.

Notice how evaluating f at a point is the same as evaluating f at its corresponding point.
1

We need to do this process twice more, for the other two edges of the triangle.
Along the left edge, along the line y = 3x + 1 , we substitute 3x + 1 in for y in f (x, y):
2 2 2
f (x, y) = f (x, 3x + 1) = x − (3x + 1 ) + 5 = −8 x − 6x + 4 = f2 (x). (12.8.10)

We want the maximum and minimum values of f on the interval 2 [−1, 0] , so we evaluate f2 at its critical points and the
endpoints of the interval. We find the critical points:

f (x) = −16x − 6 = 0 ⇒ x = −3/8. (12.8.11)
2

Evaluate f at its critical point and the endpoints of [−1, 0]:


2

f2 (−1) = 2 ⇒ f (−1, −2) = 2

f2 (−3/8) = 41/8 = 5.125 ⇒ f (−3/8, −0.125) = 5.125

f2 (0) = 1 ⇒ f (0, 1) = 4.

Finally, we evaluate f along the right edge of the triangle, where y = −3/2x + 1 .

2 2
5 2
f (x, y) = f (x, −3/2x + 1) = x − (−3/2x + 1 ) +5 = − x + 3x + 4 = f3 (x). (12.8.12)
4

The critical points of f


3 (x) are:

f (x) = 0 ⇒ x = 6/5 = 1.2. (12.8.13)
3

12.8.6 https://math.libretexts.org/@go/page/7580
We evaluate f at this critical point and at the endpoints of the interval [0, 2]:
3

f3 (0) = 4 ⇒ f (0, 1) = 4

f3 (1.2) = 5.8 ⇒ f (1.2, −0.8) = 5.8

f3 (2) = 5 ⇒ f (2, −2) = 5.

One last point to test: the critical point of f , (0, 0). We find f (0, 0) = 5 .

Figure 12.32: The surface of f along with important points along the boundary of S and the interior.
We have evaluated f at a total of 7 different places, all shown in Figure 12.32. We checked each vertex of the triangle twice, as
each showed up as the endpoint of an interval twice. Of all the z -values found, the maximum is 5.8, found at (1.2, −0.8); the
minimum is 1, found at (0, −2).

This portion of the text is entitled "Constrained Optimization'' because we want to optimize a function (i.e., find its maximum
and/or minimum values) subject to a constraint -- limits on which input points are considered. In the previous example, we
constrained ourselves by considering a function only within the boundary of a triangle. This was largely arbitrary; the function and
the boundary were chosen just as an example, with no real "meaning'' behind the function or the chosen constraint.
However, solving constrained optimization problems is a very important topic in applied mathematics. The techniques developed
here are the basis for solving larger problems, where more than two variables are involved.
We illustrate the technique once more with a classic problem.

Example 12.8.7: Constrained Optimization

The U.S. Postal Service states that the girth+length of Standard Post Package must not exceed 130''. Given a rectangular box,
the "length'' is the longest side, and the "girth'' is twice the width+height.
Given a rectangular box where the width and height are equal, what are the dimensions of the box that give the maximum
volume subject to the constraint of the size of a Standard Post Package?
Solution
Let w , h and
denote the width, height and length of a rectangular box; we assume here that w = h . The girth is then

2(w + h) = 4w . The volume of the box is V (w, ℓ) = whℓ = w ℓ . We wish to maximize this volume subject to the constraint
2

4w + ℓ ≤ 130 , or ℓ ≤ 130 − 4w . (Common sense also indicates that ℓ > 0, w > 0 .)

We begin by finding the critical values of V . We find that V w = 2wℓ and Vℓ =w


2
; these are simultaneously 0 only at (0, 0).
This gives a volume of 0, so we can ignore this critical point.
We now consider the volume along the constraint ℓ = 130 − 4w. Along this line, we have:
2 2 3
V (w, ℓ) = V (w, 130 − 4w) = w (130 − 4w) = 130 w − 4w = V1 (w). (12.8.14)

The constraint is applicable on the w-interval [0, 32.5] as indicated in the figure. Thus we want to maximize V on [0, 32.5]. 1

Finding the critical values of V , we take the derivative and set it equal to 0:
1

12.8.7 https://math.libretexts.org/@go/page/7580
2
260

V (w) = 260w − 12 w =0 ⇒ w(260 − 12w) = 0 ⇒ w = 0, ≈ 21.67. (12.8.15)
1
12

We found two critical values: when w = 0 and when w = 21.67. We again ignore the w = 0 solution; the maximum volume,
subject to the constraint, comes at w = h = 21.67 , ℓ = 130 − 4(21.6) = 43.33. This gives a volume of
V (21.67, 43.33) ≈ 19, 408 in . 3

Figure 12.33: Graphing the volume of a box with girth 4w and length l, subject to a size constraint.
The volume function V (w, ℓ) is shown in Figure 12.33 along with the constraint ℓ = 130 − 4w . As done previously, the
constraint is drawn dashed in the xy-plane and also along the surface of the function. The point where the volume is
maximized is indicated.

It is hard to overemphasize the importance of optimization. In "the real world,'' we routinely seek to make something better. By
expressing the something as a mathematical function, "making something better'' means "optimize some function.''
The techniques shown here are only the beginning of an incredibly important field. Many functions that we seek to optimize are
incredibly complex, making the step of "find the gradient and set it equal to 0⃗'' highly nontrivial. Mastery of the principles here is
key to being able to tackle these more complicated problems.

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 12.8: Extreme Values is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et
al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

12.8.8 https://math.libretexts.org/@go/page/7580
12.E: Applications of Functions of Several Variables (Exercises)
12.1: Introduction to Multivariable Functions
Terms and Concepts
1. Give two examples (other than those given in the text) of "real world" functions that require more than one input.
2. The graph of a function of two variables is a _______.
3. Most people are familiar with the concept of level curves in the context of ______ maps.
4. T/F: Along a level curve, the output of a function does not change.
5. The analogue of a level curve for functions of three variables is a level _______.
6. What does it mean when level curves are close together? Far apart?

Problems
In Exercises 7-14, give the domain and range of the multivariable function.
7. f (x, y) = x 2
+y
2
+2

8. f (x, y) = x + 2y
9. f (x, y) = x − 2y
10. f (x, y) = 1

x+2y

11. f (x, y) = 1

x2 +y 2 +1

12. f (x, y) = sin x cos y


−−−−−−−− −
13. f (x, y) = √9 − x 2
− y2

14. f (x, y) = 1

√x2 +y 2 −9

In Exercises 15-22, describe in words and sketch the level curves for the function and given c values.
15. f (x, y) = 3x − 2y; c = −2, 0, 2

16. f (x, y) = x 2 2
− y ; c = −1, 0, 1

17. f (x, y) = x − y 2
; c = −2, 0, 2

2 2
1−x −y
18. f (x, y) = 2y−2x
; c = −2, 0, 2

2x−2y
19. f (x, y) = 2
x +y
2
+1
; c = −1, 0, 1

3
y−x −1
20. f (x, y) = x
; c = −3, −1, 0, 1, 3

−− −−−−−
21. 2 2
f (x, y) = √x + 4 y ; c = 1, 2, 3, 4

22. f (x, y) = x 2 2
+ 4 y ; c = 1, 2, 3, 4

In Exercises 23-26, give the domain and range of the functions of three variables.
23. f (x, y, z) = x+2y−4z
x

24. f (x, y, z) = 1−x −y


2
1
2
−z
2

−−−−−−−− −
25. 2
f (x, y, z) = √z − x + y
2

26. f (x, y, z) = z 2
sin x cos y

In Exercises 27-30, describe the level surfaces of the given functions of three variables.
27. f (x, y, z) = x 2
+y
2
+z
2

12.E.1 https://math.libretexts.org/@go/page/10003
28. f (x, y, z) = z − x 2
+y
2

2 2
x +y
29. f (x, y, z) = z

30. f (x, y, z) = x−y


z

31. Compare the level curves of Exercises 21 and 22. How are they similar, and how are they different? Each surface is a quadric
surface; describe how the level curves are consistent with what we know about each surface.

12.2: Limits and Continuity of Multivariable Functions


Terms and Concepts
1. Describe in your own words the difference between boundary and interior point of a set.
2. Use your own words to describe (informally) what lim f (x, y) = 17 means.
(x,y)→(1,2)

3. Give an example of a closed, bounded set.


4. Give an example of a closed, unbounded set.
5. Give an example of a open, bounded set.
6. Give an example of a open, unbounded set.

Problems
In Exercises 7-10, a set S is given.
(a) Give one boundary point and one interior point, when possible, of S.
(b) State whether S is open, closed, or neither.
(c) State whether S is bounded or unbounded.
2 2
(x−1) (y−3)
7. S = {(x, y)∣∣ 4
+
9
≤ 1}

8. S = {(x, y)∣∣y ≠ x 2
}

9. S = {(x, y)∣∣x 2
+y
2
= 1}

10. S = {(x, y)∣∣y > sin x}

In Exercises 11-14:
(a) Find the domain D of the given function.
(b) State whether D is open or closed set.
(c) State whether D is bounded or unbounded.
−−−−−−−− −
11. f (x, y) = √9 − x 2
−y
2

−−−−−
12. f (x, y) = √y − x 2

13. f (x, y) = 1

√y−x2

2 2
x −y
14. f (x, y) = x +y
2 2

In Exercises 15-20, a limit is given. Evaluate the limit along the paths given, then state why these results show the given
limit does not exist.
2 2
x −y
15. lim 2
x +y
2
(x,y)→(0,0)

(a) Along the path y = 0 .


(b) Along the path x = 0 .
x+y
16. lim
x−y
(x,y)→(0,0)

(a) Along the path y = mx .

12.E.2 https://math.libretexts.org/@go/page/10003
2
xy−y
17. lim
y
2
+x
(x,y)→(0,0)

(a) Along the path y = mx .


(b) Along the path x = 0 .
2
sin( x )
18. lim
y
(x,y)→(0,0)

(a) Along the path y = mx .


(b) Along the path y = x . 2

x+y−3
19. lim
x −1
2
(x,y)→(1,2)

(a) Along the path y = 2 .


(b) Along the path y = x + 1 .
20. lim
sin x

cos y
(x,y)→(π,π/2)

(a) Along the path x = π .


(b) Along the path y = x − π/2 .

12.3: Partial Derivatives


Terms and Concepts
1. What is the difference between a constant and a coefficient?
2. Given a function z = f (x, y) , explain in your own words how to compute f . x

3. In the mixed partial fraction f , which is computed first, f or f ?


xy x y

2
∂ f
4. In the mixed partial fraction ∂x∂y
, which is computed first, f or f .
x y

Problems
In Exercises 5-8, evaluate f x (x, y) and f y (x, y) at the indicated point.
5. f (x, y) = x 2
y − x + 2y + 3 at (1, 2)

6. f (x, y) = x 3
− 3x + y
2
− 6y at (−1, 3)

7. f (x, y) = sin y cos x at (π/3, π/3)


8. f (x, y) = ln(xy) at (−2, −3)
In Exercises 9-26, find f x, fy , fxx , fyy , fxy and fyx .
9. f (x, y) = x 2
y + 3x
2
+ 4y − 5

10. f (x, y) = y 3
+ 3x y
2 2
+ 3x y + x
3

11. f (x, y) = x

12. f (x, y) = 4

xy

13. f (x, y) = e
2 2
x +y

14. f (x, y) = e x+2y

15. f (x, y) = sin x cos y


16. f (x, y) = (x + y) 3

17. f (x, y) = cos(5x y 3


)

18. f (x, y) = sin(5x 2


+ 2y )
3

−−−−−−−
19. f (x, y) = √4x y 2
+1

20. f (x, y) = (2x + 5y)√y

12.E.3 https://math.libretexts.org/@go/page/10003
21. f (x, y) = 1

x2 +y 2 +1

22. f (x, y) = 5x − 17y


23. f (x, y) = 3x 2
+1

24. f (x, y) = ln(x 2


+ y)

25. f (x, y) = ln x

4y

26. f (x, y) = 5e x
sin y + 9

In Exercises 27-30, form a function z = f (x, y) such that f and f match those given.
x y

27. fx = sin y + 1, fy = x cos y

28. fx = x + y, fy = x + y

29. fx = 6xy − 4 y ,
2
fy = 3 x
2
− 8xy + 2

2y
30. fx =
2x
2
x +y
2
, fy = 2
x +y
2

In Exercises 31-34, find f x, fy , fz , fyz and f .


zy

31. f (x, y, z) = x 2
e
2y−3z

32. f (x, y, z) = x 3
y
2 3
+x z+y z
2

33. f (x, y, z) = 7y
3x
2
z

34. f (x, y, z) = ln(xyz)

12.4: Differentiability and the Total Differential


Terms and Concepts
1. T/F: If f (x, y) is differentiable on S, the f is continuous on S.
2. T/F: If f and f are continuous on S, then f is differentiable on S.
x y

3. T/F: If z = f (x, y) is differentiable, then the change in z over small changes dx and dy in x and y is approximately dz.
4. Finish the sentence: "The new z-value is approximately the old z-value plus the approximate ______."

Problems
In Exercises 5-8, find the total differential dz.
5. z = x sin y + x 2

6. z = (2x 2
+ 3y )
2

7. z = 5x − 7y
8. z = xe x+y

In Exercises 9-12, a function z = f (x, y) is given. Give the indicated approximation using the total differential.
−−−−−
9. f (x, y) = √x 2
+y . Approximate f (2, 95, 7.1)knowing f (3, 7) = 4 .
10. f (x, y) = sin x cos y. Approximate f (0.1, −0.1)knowing f (0, 0) = 0 .
11. f (x, y) = x 2
y − xy
2
. Approximate f (2.04, 3.06)knowing f (2, 3) = −6.
12. f (x, y) = ln(x − y) . Approximate f (5.1, 3.98)knowing f (5, 4) = 0 .
Exercises 13-16 ask a variety of questions dealing with approximating error and sensitivity analysis.
13. A cylindrical storage tank is to be 2ft tall with a radius of 1ft. Is the volume of the tank more sensitive to changes in the radius
or the height?

12.E.4 https://math.libretexts.org/@go/page/10003
14. Projectile Motion: The x-value of an object moving under the principles of projectile motion is x(θ, v , t) = (v cos θ)t . A o o

particular projectile is fired with an initial velocity of v = 250 ft/s and an angle of elevation of θ = 60 . It travels a distance of
o

375ft in 3 seconds. Is the projectile more sensitive to errors in initial speed or angle of elevations?
15. The length l of a long wall is to be approximated. The angle θ , as shown in the diagram (not to scale), is measured to be 85 , ∘

and the distance x is measured to be 30 . Assume that the triangle formed is a right triangle.

Is the measurement of the length of l more sensitive to errors in the measurement of x or ln θ.


16. It is "common sense" that it is far better to measure a long distance with a long measuring tape rather than a short one. A
measured distance D can be viewed as the product of the length l of a measuring tape times the number n of times it was used. for
instance, using a 3' tape 10 times gives a length of 30 . To measure the same distance with a 12 tape, we would use the tape 2.5
′ ′

times. (i.e., 30 = 12 × 2.5 .) Thus D = nl .


Suppose each time a measurement is taken with the tape of the recorded distance within 1/16 of the actual distance, (i.e., ′′

dl = 1/ 16 ≈ 0.005 ft). Using differentials, show why common sense proves correct in that it is better to use a long tape to
′′

measure long distances.


In Exercises 17-18, find the total differential dw.
17. w = x 2
yz
3

18. w = e x
sin y ln z

In Exercises 19-22, use the information provided and the total differential to make the given approximation.
19. f (3, 1) = 7, fx (3, 1) = 9, fy (3, 1) = −2. Approximate f (3.05, 0.9).
20. f (−4, 2) = 13, fx (−4, 2) = 2.6, fy (−4, 2) = 5.1. Approximate f (−4.12, 2.07).
21. f (2, 4, 5) = −1, fx (2, 4, 5) = 2, fy (2, 4, 5) = −3, fz (2, 4, 5) = 3.7 Approximate f (2.5, 4.1, 4.8).
22. f (3, 3, 3) = 5, fx (3, 3, 3) = 2, fy (3, 3, 3) = 0, fz (3, 3, 3) = −2 Approximate f (3.1, 3.1, 3.1).

12.5: The Multivariable Chain Rule


Terms and Concepts
1. Let a level curve of z = f (x, y) be described by x = g(t), y = h(t) . Explain why dz

dt
=0 .
2. Fill in the blank: The single variable Chain Rule states d

dx
(f (g(x))) = f

(g(x)) ⋅ _________.
df ∂f dy
3. Fill in the blank: The Multivariable Chain Rule states dt
=
∂x
⋅ _________ + __________⋅ dt
.
4. If z = f (x, y) , where x = g(t) and y = h(t) , we can substitute and write z as an explicit function of t.
T/F: Using the Multivariable Chain Rule to find is sometimes easier than first substituting and then taking the derivative.
dz

dt

5. T/F: The Multivariable Chain Rule is only useful when all the related functions are known explicitly.
6. The Multivariable Chain Rule allows us to compute implicit derivatives easily by just computing two ______ derivatives.

Problems
In Exercises 7-12, functions z = f (x, y), x = g(t) and y = h(t) are given.
(a) Use the Multivariable Chain Rule to compute . dz

dt

(b) Evaluate at the indicated t-value.


dz

dt

7. z = 3x + 4y, x =t ,
2
y = 2t; t =1

8. z = x 2
−y ,
2
x = t, y =t
2
− 1; t =1

9. z = 5x + 2y, x = 2 cos t + 1, y = sin t − 3; t = π/4

10. z = y
x
2
+1
, x = cos t, y = sin t; t = π/2

11. z = x 2
+ 2y ,
2
x = sin t, y = 3 sin t; t = π/4

12.E.5 https://math.libretexts.org/@go/page/10003
12. z = cos x sin y, x = πt, y = 2πt + π/2; t =3

In Exercises 13-18, functions z = f (x, y), x = g(t) and y = h(t) are given. Find the values of t where dz

dt
=0 . Note: these
are the same surfaces/curves as found in Exercises 7-12.
13. z = 3x + 4y, x =t ,
2
y = 2t

14. z = x 2
−y ,
2
x = t, y =t
2
−1

15. z = 5x + 2y, x = 2 cos t + 1, y = sin t − 3

16. z = y
x
2
+1
, x = cos t, y = sin t

17. z = x 2
+ 2y ,
2
x = sin t, y = 3 sin t

18. z = cos x sin y, x = πt, y = 2πt + π/2

In Exercises 19-22, functions z = f (x, y), x = g(s, t) and y = h(s, t) are given.
(a) Use the Multivariable Chain Rule to compute and \(\frac{\partial z}{\partial t}\). ∂z

∂s

(b) Evaluate ∂z

∂s
and ∂z

∂t
at the indicated s and t values.
19. z = x 2
y, x = s − t, y = 2s + 4t; s = 1, t = 0

20. z = cos(πx + π

2
y), x = st ,
2
y = s t;
2
s = 1, t = 1

21. z = x 2
+y ,
2
x = s cos t, y = s sin t; s = 2, t = π/4

22. z = e
2 2
−( x +y ) 2
, x = t, y = st ; s = 1, t = 1

dy
In Exercises 23-26, find dx
using Implicit Differentiation and Theorem 109.
23. x 2
tan y = 50

24. (3x 2
+ 2y )
3 4
=2

2
x +y
25. x+y 2
= 17

26. ln(x 2
+ xy + y ) = 1
2

In Exercises 27-30, find dz

dt
, or ∂z

∂s
and ∂z

∂t
, using the supplied information.
dy
27. ∂z

∂x
= 2,
∂z

∂y
= 1,
dx

dt
= 4,
dt
= −5

dy
28. ∂z

∂x
= 1,
∂z

∂y
= −3,
dx

dt
= 6,
dt
=2

∂y ∂y
29. ∂z

∂x
= −4,
∂z

∂y
= 9,
∂x

∂s
= 5,
∂x

∂t
= 7,
∂s
= −2,
∂t
=6

∂y ∂y
30. ∂z

∂x
= 2,
∂z

∂y
= 1,
∂x

∂s
= −2,
∂x

∂t
= 3,
∂s
= 2,
∂t
= −1

12.6: Directional Derivatives


Terms and Concepts
1. What is the difference between a directional derivative and a partial derivative?
2. For what u⃗ is D u⃗
f = fx ?
3. For what u⃗ is D u⃗
f = fy ?
4. The gradient is _______ to level curves.
5. The gradient points in the direction of _______ increase.
6. It is generally more informative to view the directional derivative not as the result of a limit, but rather as the result of a
________ product.

12.E.6 https://math.libretexts.org/@go/page/10003
Problems
In Exercises 7-12, a function z = f (x, y) is given. Find ∇f .
7. f (x, y) = −x 2
y + xy
2
+ xy

8. f (x, y) = sin x cos y


9. f (x, y) = 1

x2 +y 2 +1

10. f (x, y) = −4x + 3y


11. f (x, y) = x 2
+ 2y
3
− xy − 7x

12. f (x, y) = x 2
y
3
− 2x

In Exercises 13-18, a function z = f (x, y) and a point P are given. Find the directional derivative of f in the indicated
directions. Note: these are the same functions as in Exercises 7-12.
13. f (x, y) = −x y + x y + xy, P = (2, 1)
2 2

(a) In the direction of v ⃗ = ⟨3, 4⟩


(b) In the direction toward the point Q = (−1, 1)
14. f (x, y) = sin x cos y, P = ( , ) π

4
π

(a) In the direction of v⃗ = ⟨1, 1⟩


(b) In the direction toward the point Q = (0, 0)
15. f (x, y) = x2 +y 2 +1
1
, P = (1, 1)

(a) In the direction of v ⃗ = ⟨1, −1⟩


(b) In the direction toward the point Q = (−2, −2)
16. f (x, y) = −4x + 3y, P = (5, 2)
(a) In the direction of v ⃗ = ⟨3, 1⟩
(b) In the direction toward the point Q = (2, 7)
17. f (x, y) = x + 2y − xy − 7x, P = (4, 1)
2 2

(a) In the direction of v ⃗ = ⟨−2, 5⟩


(b) In the direction toward the point Q = (4, 0)
18. f (x, y) = x y − 2x, P = (1, 1)
2 3

(a) In the direction of v ⃗ = ⟨3, 3⟩


(b) In the direction toward the point Q = (1, 2)
In Exercises 19-24, a function z = f (x, y) and a point P are given.
(a) Find the direction of maximal increase of f at P .
(b) What is the maximal value of D f at P ? u⃗

(c) Find the direction of minimal increase of f at P .


(d) Give a direction u⃗ such that D f = 0 at P . u⃗

Note: these are the same functions and points as in Exercises 13 through 18.
19. f (x, y) = −x 2
y + xy
2
+ xy, P = (2, 1)

20. f (x, y) = sin x cos y, P =(


π

4
,
π

3
)

21. f (x, y) = 2
x +y
1
2
+1
, P = (1, 1)

22. f (x, y) = −4x + 3y, P = (5, 4)

23. f (x, y) = x 2
+ 2y
2
− xy − 7x, P = (4, 1)

24. f (x, y) = x 2
y
3
− 2x, P = (1, 1)

In Exercises 25-28, a function w = F (x, y, z), a vector v ⃗ and a point P are given.
(a) Find ∇F (x, y, z).
(b) Find D F at P.
u⃗

12.E.7 https://math.libretexts.org/@go/page/10003
25. f (x, y) = 3x 2
z
3 2
+ 4xy − 3 z , v ⃗ = ⟨1, 1, 1⟩, P = (3, 2, 1)

26. f (x, y) = sin(x) cos(y)e z


, v⃗ = ⟨2, 2, 1⟩, P = (0, 0, 0)

27. f (x, y) = x 2
y
2 2 2
− y z , v ⃗ = ⟨−1, 7, 3⟩, P = (1, 0, −1)

28. f (x, y) = 2
x +y
2
2
+z
2
, v ⃗ = ⟨1, 1, −2⟩, P = (1, 1, 1)

12.7: Tangent Lines, Normal Lines, and Tangent Planes


Terms and Concepts
1. Explain how the vector v ⃗ = ⟨1, 0, 3⟩ can be thought of as having a "slope" of 3.
2. Explain how the vector v ⃗ = ⟨0.6, 0.8, −2⟩ can be thought of as having a "slope" of -2.
3. T/F: Let z = f (x, y) be differentiable at P. If n⃗ is normal vector to the tangent plane of f at P , then n⃗ is orthogonal to f and x

f at P.
y

4. Explain in your own words why we do not refer to the tangent line to a surface at a point, but rather to directional tangent lines to
a surface at a point.

Problems
In Exercises 5-8, a function z = f (x, y) , a vector v ⃗ and a point P are given. Give the parametric equations of the following
directional tangent lines to f at P :
(a) l (t)
x

(b) l (t)
y

(c) l , where u⃗ is the unit vector in the direction of v ⃗.


u⃗

5. f (x, y) = 2x y 2
− 4x y , v⃗ = ⟨1, 3⟩, P = (2, 3) .
6. f (x, y) = 3 cos x sin y, v ⃗ = ⟨1, 2⟩, P = (π/3, π/6) .
7. f (x, y) = 3x − 5y, v ⃗ = ⟨1, 1⟩, P = (4, 2) .
8. f (x, y) = x 2
− 2x − y
2
+ 4y, v ⃗ = ⟨1, 1⟩, P = (1, 2) .
In Exercises 9-12, a function z = f (x, y) and a point P are given. Find the equation of the normal line to f at P . Note: these
are the same functions as in Exercises 5-8.
9. f (x, y) = 2x y 2
− 4x y , v ⃗ = ⟨1, 3⟩, P = (2, 3) .
10. f (x, y) = 3 cos x sin y, v ⃗ = ⟨1, 2⟩, P = (π/3, π/6) .
11. f (x, y) = 3x − 5y, v ⃗ = ⟨1, 1⟩, P = (4, 2) .
12. f (x, y) = x 2
− 2x − y
2
+ 4y, v⃗ = ⟨1, 1⟩, P = (1, 2) .
In Exercises 13-16, a function z = f (x, y) and a point P are given. Find the two points that are 2 units from the surface f at
P . Note: these are the same functions as in Exercises 5-8.

13. f (x, y) = 2x y 2
− 4x y , v ⃗ = ⟨1, 3⟩, P = (2, 3) .
14. f (x, y) = 3 cos x sin y, v ⃗ = ⟨1, 2⟩, P = (π/3, π/6) .
15. f (x, y) = 3x − 5y, v ⃗ = ⟨1, 1⟩, P = (4, 2) .
16. f (x, y) = x 2
− 2x − y
2
+ 4y, v⃗ = ⟨1, 1⟩, P = (1, 2) .
In Exercises 17-20, a function z = f (x, y) and a point P are given. Find the equation of the tangent plane to f at P . Note:
these are the same functions as in Exercises 5-8.
17. f (x, y) = 2x y 2
− 4x y , v ⃗ = ⟨1, 3⟩, P = (2, 3) .
18. f (x, y) = 3 cos x sin y, v ⃗ = ⟨1, 2⟩, P = (π/3, π/6) .
19. f (x, y) = 3x − 5y, v ⃗ = ⟨1, 1⟩, P = (4, 2) .
20. f (x, y) = x 2
− 2x − y
2
+ 4y, v⃗ = ⟨1, 1⟩, P = (1, 2) .

12.E.8 https://math.libretexts.org/@go/page/10003
In Exercises 21-24, an implicitly defined function of x, y and z is given along with a point P that lies on the surface. Use the
gradient ∇F to:
(a) find the equation of the normal line to the surface at P, and
(b) find the equation of the plane tangent to the surface at P.
2
y
2 2
– –
21. x

8
+
4
+
z

16
= 1, at P = (1, √2, √6)

2 y
2

22. z 2

x

4

9
= 0, at P = (4, −3, √5)

23. x y 2
− xz
2
= 0, at P = (2, 1, −1)

24. sin(xy) + cos)yz) = 0, at P = (2, π/12, 4)

12.8: Extreme Values


Terms and Concepts
1. T/F: Theorem 114 states that if f has a critical point at P , then f has a relative extrema at P .
2. T/F: A point P is a critical point of f if f and f are both 0 at P . x y

3. T/F: A point P is a critical point of f if f or f are undefined at P . x y

4. Explain what it means to "solve a constrained optimization" problem.

Problems
Exercises 5-14, find the critical points of the given function. Use the Second Derivative Test to determine if each critical
point corresponds to a relative maximum, minimum, or saddle point.
5. f (x, y) = 1

2
x
2
+ 2y
2
− 8y + 4x

6. f (x, y) = x 2
+ 4x + y
2
− 9y + 3xy

7. f (x, y) = x 2
+ 3y
2
− 6y + 4xy

8. f (x, y) = x +y
2
1
2
+1

9. f (x, y) = x 2
+y
3
− 3y + 1

10. f (x, y) = 1

3
x
3
−x +
1

3
y
3
− 4y

11. f (x, y) = x 2
y
2

12. f (x, y) = x 4
− 2x
2
+y
3
− 27y − 15

−−−−− −−− −− − −−−−


13. 2
f (x, y) = √16 − (x − 3 ) − y
2

−−−−−−
14. f (x, y) = √x 2
+y
2

In Exercises 15-18, find the absolute maximum and minimum of the function subject to the given constraint.
15. f (x, y) = x 2
+y
2
+y = 1 , constrained to the triangles with vertices 0, 1), (−1, 1), and (1, −1) .
16. f (x, y) = 5x − 7y , constrained to the region bounded by y = x 2
and y = 1 .
17. f (x, y) = x 2
+ 2x + y
2
+ 2y , constrained to the region bounded by the circle x 2
+y
2
=4 .
18. f (x, y) = 3y − 2x , constrained to the region bounded by the parabola y = x
2 2
+x −1 and the line y = x .

12.E: Applications of Functions of Several Variables (Exercises) is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or
curated by LibreTexts.

12.E.9 https://math.libretexts.org/@go/page/10003
CHAPTER OVERVIEW
13: Multiple Integration
In this chapter we apply techniques of integral calculus to multivariable functions. In Chapter 5 we learned how the definite
integral of a single variable function gave us "area under the curve." In this chapter we will see that integration applied to a
multivariable function gives us "volume under a surface." And just as we learned applications of integration beyond finding areas,
we will find applications of integration in this chapter beyond finding volume.
13.1: Iterated Integrals and Area
13.2: Double Integration and Volume
13.3: Double Integration with Polar Coordinates
13.4: Center of Mass
13.5: Surface Area
13.6: Volume Between Surfaces and Triple Integration
13.E: Applications of Multiple Integration (Exercises)

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 13: Multiple Integration is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman
et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

1
13.1: Iterated Integrals and Area
In the previous chapter we found that we could differentiate functions of several variables with respect to one variable, while
treating all the other variables as constants or coefficients. We can integrate functions of several variables in a similar way. For
instance, if we are told that f (x, y) = 2xy, we can treat y as staying constant and integrate to obtain f (x, y):
x

f (x, y) = ∫ fx (x, y) dx

=∫ 2xy dx

2
= x y + C.

Make a careful note about the constant of integration, C . This "constant'' is something with a derivative of 0 with respect to x, so it
could be any expression that contains only constants and functions of y . For instance, if f (x, y) = x y + sin y + y + 17 , then 2 3

f (x, y) = 2xy. To signify that C is actually a function of y , we write:


x

2
f (x, y) = ∫ fx (x, y) dx = x y + C (y). (13.1.1)

Using this process we can even evaluate definite integrals.

Example 13.1.1: Integrating functions of more than one variable


2y

Evaluate the integral ∫ 2xy dx.


1

Solution
We find the indefinite integral as before, then apply the Fundamental Theorem of Calculus to evaluate the definite integral:
2y
2y
2 ∣
∫ 2xy dx = x y

1
1

2 2
= (2y ) y − (1 ) y
3
= 4y − y.

We can also integrate with respect to y . In general,


h2 (y)
h2 (y)

∫ fx (x, y) dx = f (x, y) = f (h2 (y), y) − f (h1 (y), y), (13.1.2)

h1 (y)
h1 (y)

and
g2 (x)
g2 (x)

∫ fy (x, y) dy = f (x, y) = f (x, g2 (x)) − f (x, g1 (x)). (13.1.3)

g1 (x)
g1 (x)

Note that when integrating with respect to x, the bounds are functions of y (of the form x = h (y) and x = h (y) ) and the final 1 2

result is also a function of y . When integrating with respect to y , the bounds are functions of x (of the form y = g (x) and 1

y = g (x) ) and the final result is a function of x. Another example will help us understand this.
2

Example 13.1.2: Integrating functions of more than one variable


x

Evaluate ∫ 3
(5 x y
−3
+ 6 y ) dy
2
.
1

Solution
We consider x as staying constant and integrate with respect to y :

13.1.1 https://math.libretexts.org/@go/page/4234
x 3 −2 3 x
5x y 6y ∣
3 −3 2
∫ (5 x y + 6 y ) dy = ( + )∣
1
−2 3 ∣
1

5 3 −2 3
5 3
= (− x x + 2 x ) − (− x + 2)
2 2

9 5
3
= x − x − 2.
2 2

Note how the bounds of the integral are from y = 1 to y = x and that the final answer is a function of x.

In the previous example, we integrated a function with respect to y and ended up with a function of x. We can integrate this as
well. This process is known as iterated integration, or multiple integration.

Example 13.1.3: Integrating an integral


2 x

Evaluate ∫ (∫
3
(5 x y
−3 2
+ 6 y ) dy) dx.
1 1

Solution
We follow a standard "order of operations'' and perform the operations inside parentheses first (which is the integral evaluated
in Example 13.1.2.)
2 x 2 3 −2 3 x
5x y 6y ∣
3 −3 2
∫ (∫ (5 x y + 6 y ) dy) dx =∫ ([ + ] ∣ ) dx
1 1 1
−2 3 ∣
1

2
9 5
3
=∫ ( x − x − 2) dx
1 2 2
2
9 5 ∣
4 2
=( x − x − 2x) ∣
8 4 ∣
1

89
= .
8

Note how the bounds of x were x = 1 to x = 2 and the final result was a number.

The previous example showed how we could perform something called an iterated integral; we do not yet know why we would be
interested in doing so nor what the result, such as the number 89/8, means. Before we investigate these questions, we offer some
definitions.

Definition: Iterated Integration


Iterated integration is the process of repeatedly integrating the results of previous integrations. Integrating one integral is
denoted as follows.
Let a , b , c and d be numbers and let g 1 (x) ,g 2 (x) ,h
1 (y) and h 2 (y) be functions of x and y , respectively. Then:
d h2 (y) d h2 (y)

1. ∫ ∫ f (x, y) dx dy = ∫ (∫ f (x, y) dx) dy.


c h1 (y) c h1 (y)

b g2 (x) b g2 (x)

2. ∫ ∫ f (x, y) dy dx = ∫ (∫ f (x, y) dy) dx.


a g (x) a g (x)
1 1

Again make note of the bounds of these iterated integrals.


d h2 (y)

With ∫ ∫ f (x, y) dx dy x , varies from h1 (y) to h2 (y) , whereas y varies from c to d . That is, the bounds of x are curves,
c h1 (y)

the curves x = h (y) and x = h (y) , whereas the bounds of y are constants, y = c and y = d . It is useful to remember that when
1 2

setting up and evaluating such iterated integrals, we integrate "from curve to curve, then from point to point.''
We now begin to investigate why we are interested in iterated integrals and what they mean.

13.1.2 https://math.libretexts.org/@go/page/4234
Area of a plane region
Consider the plane region R bounded by a ≤ x ≤ b and g1 (x) ≤ y ≤ g2 (x) , shown in Figure 13.1.1 . We learned in Section 7.1
(in Calculus I) that the area of R is given by
b

∫ (g2 (x) − g1 (x)) dx. (13.1.4)


a

Figure 13.1.1 : Calculating the area of a plane region R with an iterated integral.
We can view the expression (g 2 (x) − g1 (x)) as
g (x) g (x)
2 2

(g2 (x) − g1 (x)) = ∫ 1 dy = ∫ dy,


g1 (x) g1 (x)

meaning we can express the area of R as an iterated integral:


b b g (x) b g (x)
2 2

area of R = ∫ (g2 (x) − g1 (x)) dx = ∫ (∫ dy) dx = ∫ ∫ dy dx. (13.1.5)


a a g1 (x) a g1 (x)

In short: a certain iterated integral can be viewed as giving the area of a plane region.
A region R could also be defined by c ≤ y ≤ d and h 1 (y) ≤ x ≤ h2 (y) , as shown in Figure 13.1.2. Using a process similar to that
above, we have
d h2 (y)

the area of R = ∫ ∫ dx dy. (13.1.6)


c h1 (y)

Figure 13.1.2 : Calculating the area of a plane region R with an iterated integral.
We state this formally in a theorem.

THEOREM 13.1.1: Area of a plane region

1. Let R be a plane region bounded by a ≤ x ≤ b and g 1 (x) ≤ y ≤ g2 (x) , where g and g are continuous functions on
1 2

[a, b]. The area A of R is

b g2 (x)

A =∫ ∫ dy dx. (13.1.7)
a g1 (x)

13.1.3 https://math.libretexts.org/@go/page/4234
2. Let R be a plane region bounded by c ≤ y ≤ d and h 1 (y) ≤ x ≤ h2 (y) , where h and h are continuous functions on
1 2

[c, d]. The area A of R is

d h2 (y)

A =∫ ∫ dx dy. (13.1.8)
c h1 (y)

The following examples should help us understand this theorem.

Example 13.1.4: Area of a rectangle

Find the area A of the rectangle with corners (−1, 1) and (3, 3), as shown in Figure 13.1.3.

Figure 13.1.3 : Calculating the area of a rectangle with an iterated integral in Example 13.1.4 .
Solution
Multiple integration is obviously overkill in this situation, but we proceed to establish its use.
The region R is bounded by x = −1 , x = 3 , y = 1 and y = 3 . Choosing to integrate with respect to y first, we have
3 3 3 3
3 3
∣ ∣
A =∫ ∫ 1 dy dx = ∫ (y ) dx = ∫ 2 dx = 2x = 8.
∣ ∣
1 −1
−1 1 −1 −1

We could also integrate with respect to x first, giving:


3 3 3 3
3 3
∣ ∣
A =∫ ∫ 1 dx dy = ∫ (x ) dy = ∫ 4 dy = 4y = 8.
∣−1 ∣1
1 −1 1 1

Clearly there are simpler ways to find this area, but it is interesting to note that this method works.

Example 13.1.5: Area of a triangle

Find the area A of the triangle with vertices at (1, 1), (3, 1) and (5, 5), as shown in Figure 13.1.4.

Figure 13.1.4 : Calculating the area of a triangle with iterated integrals in Example 13.1.5 .
Solution

13.1.4 https://math.libretexts.org/@go/page/4234
The triangle is bounded by the lines as shown in the figure. Choosing to integrate with respect to x first gives that x is bounded
y+5
by x = y to x = 2
, while y is bounded by y = 1 to y = 5 . (Recall that since x-values increase from left to right, the
leftmost curve, x = y , is the lower bound and the rightmost curve, x = (y + 5)/2 , is the upper bound.) The area is
y+5
5
2

A =∫ ∫ dx dy
1 y

y+5
5

∣ 2
=∫ (x ) dy
∣y
1

5
1 5
=∫ (− y+ ) dy
1
2 2

1 5 5
2 ∣
= (− y + y)

4 2 1

= 4.

We can also find the area by integrating with respect to y first. In this situation, though, we have two functions that act as the
lower bound for the region R , y = 1 and y = 2x − 5 . This requires us to use two iterated integrals. Note how the x-bounds are
different for each integral:
3 x 5 x

A =∫ ∫ 1 dy dx + ∫ ∫ 1 dy dx
1 1 3 2x−5

3 5
x x
∣ ∣
=∫ (y) dx + ∫ (y) dx
∣1 ∣2x−5
1 3

3 5

=∫ (x − 1) dx + ∫ ( − x + 5) dx
1 3

=2 + 2

= 4.

As expected, we get the same answer both ways.

Example 13.1.6: Area of a plane region

Find the area of the region enclosed by y = 2x and y = x , as shown in Figure 13.1.5. 2

Figure 13.1.5 : Calculating the area of a plane region with iterated integrals in Example 13.1.6 .
Solution
Once again we'll find the area of the region using both orders of integration.
Using dy dx:
2 2x 2
1 2 4
2 2 3 ∣
∫ ∫ 1 dy dx = ∫ (2x − x ) dx = (x − x ) = . (13.1.9)

0 x
2
0
3 0 3

Using dx dy:

13.1.5 https://math.libretexts.org/@go/page/4234
4 √y 4
2 1 4 4
3/2 2 ∣
∫ ∫ 1 dx dy = ∫ (√y − y/2) dy = ( y − y ) = . (13.1.10)

0 y/2 0
3 4 0 3

Changing Order of Integration


In each of the previous examples, we have been given a region R and found the bounds needed to find the area of R using both
orders of integration. We integrated using both orders of integration to demonstrate their equality.
We now approach the skill of describing a region using both orders of integration from a different perspective. Instead of starting
with a region and creating iterated integrals, we will start with an iterated integral and rewrite it in the other integration order. To do
so, we'll need to understand the region over which we are integrating.
The simplest of all cases is when both integrals are bound by constants. The region described by these bounds is a rectangle (see
Example 13.1.4), and so:
b d d b

∫ ∫ 1 dy dx = ∫ ∫ 1 dx dy. (13.1.11)
a c c a

When the inner integral's bounds are not constants, it is generally very useful to sketch the bounds to determine what the region we
are integrating over looks like. From the sketch we can then rewrite the integral with the other order of integration.
Examples will help us develop this skill.

Example 13.1.7: Changing the order of integration


6 x/3

Rewrite the iterated integral ∫ ∫ 1 dy dx with the order of integration dx dy .


0 0

Solution
We need to use the bounds of integration to determine the region we are integrating over.
The bounds tell us that y is bounded by 0 and x/3; x is bounded by 0 and 6. We plot these four curves: y = 0 , y = x/3, x =0

and x = 6 to find the region described by the bounds. Figure 13.1.6 shows these curves, indicating that R is a triangle.

Figure 13.1.6 : Sketching the region R described by the iterated integral in Example 13.1.7 .
To change the order of integration, we need to consider the curves that bound the x-values. We see that the lower bound is
2 6

x = 3y and the upper bound is x = 6 . The bounds on y are 0 to 2. Thus we can rewrite the integral as ∫ ∫ 1 dx dy.
0 3y

Example 13.1.8: Changing the order of integration


4 (y+4)/2

Change the order of integration of ∫ ∫ 1 dx dy .


0 y 2 /4

Solution
We sketch the region described by the bounds to help us change the integration order. x is bounded below and above (i.e., to
the left and right) by x = y /4 and x = (y + 4)/2 respectively, and y is bounded between 0 and 4. Graphing the previous
2

13.1.6 https://math.libretexts.org/@go/page/4234
curves, we find the region R to be that shown in Figure 13.1.7.

Figure 13.1.7 : Drawing the region determined by the bounds of integration in Example 13.1.8 .
To change the order of integration, we need to establish curves that bound y . The figure makes it clear that there are two lower
bounds for y : y = 0 on 0 ≤ x ≤ 2 , and y = 2x − 4 on 2 ≤ x ≤ 4 . Thus we need two double integrals. The upper bound for
each is y = 2√− x . Thus we have

4 (y+4)/2 2 2 √x 4 2 √x

∫ ∫ 1 dx dy = ∫ ∫ 1 dy dx + ∫ ∫ 1 dy dx.
2
0 y /4 0 0 2 2x−4

This section has introduced a new concept, the iterated integral. We developed one application for iterated integration: area
between curves. However, this is not new, for we already know how to find areas bounded by curves.
In the next section we apply iterated integration to solve problems we currently do not know how to handle. The "real" goal of this
section was not to learn a new way of computing area. Rather, our goal was to learn how to define a region in the plane using the
bounds of an iterated integral. That skill is very important in the following sections.

This page titled 13.1: Iterated Integrals and Area is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory
Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.

13.1.7 https://math.libretexts.org/@go/page/4234
13.2: Double Integration and Volume
b

The definite integral of f over [a, b], ∫ f (x) dx , was introduced as "the signed area under the curve.'' We approximated the value
a

of this area by first subdividing [a, b] into n subintervals, where the i subinterval has length Δx , and letting c be any value in
th
i i

the i subinterval. We formed rectangles that approximated part of the region under the curve with width Δx , height f (c ), and
th
i i

hence with area f (c ) Δx . Summing all the rectangle's areas gave an approximation of the definite integral, and Theorem 38 stated
i i

that
b

∫ f (x) dx = lim ∑ f (ci ) Δxi , (13.2.1)


∥Δx∥→0
a

connecting the area under the curve with sums of the areas of rectangles.
We use a similar approach in this section to find volume under a surface.
Let R be a closed, bounded region in the xy-plane and let z = f (x, y) be a continuous function defined on R . We wish to find the
signed volume under the surface of f over R . (We use the term "signed volume'' to denote that space above the xy-plane, under f ,
will have a positive volume; space above f and under the xy-plane will have a "negative'' volume, similar to the notion of signed
area used before.)
We start by partitioning R into n rectangular subregions as shown in Figure 13.2.1a. For simplicity's sake, we let all widths be Δx
and all heights be Δy. Note that the sum of the areas of the rectangles is not equal to the area of R , but rather is a close
approximation. Arbitrarily number the rectangles 1 through n , and pick a point (x , y ) in the i subregion.
i i
th

Figure 13.2.1 : Developing a method for finding signed volume under a surface.
The volume of the rectangular solid whose base is the i subregion and whose height is f (x , y ) is V = f (x , y ) Δx Δy . Such
th
i i i i i

a solid is shown in Figure 13.2.1b. Note how this rectangular solid only approximates the true volume under the surface; part of the
solid is above the surface and part is below.
For each subregion R used to approximate R , create the rectangular solid with base area Δx Δy and height f (x
i i, yi ). The sum of
all rectangular solids is
n

∑ f (xi , yi ) Δx Δy. (13.2.2)

i=1

This approximates the signed volume under f over R . As we have done before, to get a better approximation we can use more
rectangles to approximate the region R .
In general, each rectangle could have a different width Δx and height Δy , giving the i rectangle an area ΔA = Δx Δy and
j k
th
i j k

the i rectangular solid a volume of f (x , y ) ΔA . Let ||ΔA|| denote the length of the longest diagonal of all rectangles in the
th
i i i

subdivision of R ; ||ΔA|| → 0 means each rectangle's width and height are both approaching 0. If f is a continuous function, as

13.2.1 https://math.libretexts.org/@go/page/4235
n

||ΔA|| shrinks (and hence n → ∞ ) the summation ∑ f (x i, yi ) ΔAi approximates the signed volume better and better. This leads
i=1

to a definition.
b

Note: Recall that the integration symbol "∫ '' is an "elongated S,'' representing the word "sum.'' We interpreted ∫ f (x) dx as "take
a

the sum of the areas of rectangles over the interval [a, b].'' The double integral uses two integration symbols to represent a "double
sum.'' When adding up the volumes of rectangular solids over a partition of a region R , as done in Figure 13.2.1, one could first
add up the volumes across each row (one type of sum), then add these totals together (another sum), as in
n m

∑ ∑ f (xi , yj ) Δxi Δyj . (13.2.3)

j=1 i=1

One can rewrite this as


n m

∑ ( ∑ f (xi , yj ) Δxi ) Δyj . (13.2.4)

j=1 i=1

The summation inside the parenthesis indicates the sum of heights × widths, which gives an area; multiplying these areas by the
thickness Δy gives a volume. The illustration in Figure 13.2.2 relates to this understanding.
j

Definition 101: Double Integral, Signed Volume


Let z = f (x, y) be a continuous function defined over a closed region R in the xy-plane. The signed volume V under f over
R is denoted by the double integral

V =∬ f (x, y) dA. (13.2.5)


R

Alternate notations for the double integral are

∬ f (x, y) dA = ∬ f (x, y) dx dy = ∬ f (x, y) dy dx. (13.2.6)


R R R

The definition above does not state how to find the signed volume, though the notation offers a hint. We need the next two
theorems to evaluate double integrals to find volume.

theorem 118: Double Integrals and Signed Volume


Let z = f (x, y) be a continuous function defined over a closed region R in the xy-plane. Then the signed volume V under f

over R is
n

V =∬ f (x, y)dA = lim ∑ f (xi , yi ) ΔAi . (13.2.7)


||ΔA||→0
R i=1

This theorem states that we can find the exact signed volume using a limit of sums. The partition of the region R is not specified,
so any partitioning where the diagonal of each rectangle shrinks to 0 results in the same answer.
This does not offer a very satisfying way of computing area, though. Our experience has shown that evaluating the limits of sums
can be tedious. We seek a more direct method.
b
Recall Theorem 54 in Section 7.2. This stated that if A(x) gives the cross-sectional area of a solid at x, then ∫ a
A(x)dx gave the
volume of that solid over [a, b].
Consider Figure 13.2.2, where a surface z = f (x, y) is drawn over a region R . Fixing a particular x value, we can consider the
area under f over R where x has that fixed value. That area can be found with a definite integral, namely
g2 (x)

A(x) = ∫ f (x, y) dy. (13.2.8)


g (x)
1

13.2.2 https://math.libretexts.org/@go/page/4235
Remember that though the integrand contains x, we are viewing x as fixed. Also note that the bounds of integration are functions
of x: the bounds depend on the value of x.

Figure 13.2.2 : Finding volume under a surface by sweeping out a cross-sectional area.
As A(x) is a cross-sectional area function, we can find the signed volume V under f by integrating it:
b b g2 (x) b g2 (x)

V =∫ A(x) dx = ∫ (∫ f (x, y) dy) dx = ∫ ∫ f (x, y) dy dx. (13.2.9)


a a g (x) a g (x)
1 1

This gives a concrete method for finding signed volume under a surface. We could do a similar procedure where we started with y
fixed, resulting in a iterated integral with the order of integration dx dy. The following theorem states that both methods give the
same result, which is the value of the double integral. It is such an important theorem it has a name associated with it.

THEOREM 119: Fubini's Theorem


Let R be a closed, bounded region in the xy-plane and let z = f (x, y) be a continuous function on R .
1. If R is bounded by a ≤ x ≤ b and g 1 (x) ≤ y ≤ g2 (x) , where g and g are continuous functions on [a, b], then
1 2

b g2 (x)

∬ f (x, y) dA = ∫ ∫ f (x, y) dy dx. (13.2.10)


R a g1 (x)

2. If R is bounded by c ≤ y ≤ d and h 1 (y) ≤ x ≤ h2 (y) , where h and h are continuous functions on [c, d], then
1 2

d h2 (y)

∬ f (x, y) dA = ∫ ∫ f (x, y) dx dy. (13.2.11)


R c h1 (y)

Note that once again the bounds of integration follow the "curve to curve, point to point'' pattern discussed in the previous section.
In fact, one of the main points of the previous section is developing the skill of describing a region R with the bounds of an iterated
integral. Once this skill is developed, we can use double integrals to compute many quantities, not just signed volume under a
surface.

Example 13.2.1: Evaluating a double integral

Let f (x, y) = xy + e . Find the signed volume under f on the region R , which is the rectangle with corners (3, 1) and (4, 2)
y

pictured in Figure 13.2.3, using Fubini's Theorem and both orders of integration.

Solution

We wish to evaluate ∬ y
(xy + e ) dA . As R is a rectangle, the bounds are easily described as 3 ≤ x ≤ 4 and 1 ≤ y ≤ 2 .
R

Using the order dy dx:

13.2.3 https://math.libretexts.org/@go/page/4235
4 2
y y
∬ (xy + e ) dA = ∫ ∫ (xy + e ) dy dx
R 3 1

4 2
1 ∣
2 y
=∫ ([ xy + e ]∣ ) dx
3
2 ∣
1

4
3
2
=∫ ( x +e − e) dx
3 2
4
3 ∣
2 2
= ( x + (e − e)x) ∣
4 ∣
3

21
2
= +e − e ≈ 9.92.
4

Figure 13.2.3 : Finding the signed volume under a surface in Example 13.2.1.
Now we check the validity of Fubini's Theorem by using the order dx dy:
2 4
y y
∬ (xy + e ) dA = ∫ ∫ (xy + e ) dx dy
R 1 3

2 4
1 ∣
2 y
=∫ ([ x y + x e ]∣ ) dy
1
2 ∣3

2
7 y
=∫ ( y + e ) dy
1
2

2
7 ∣
2 y
= ( y + e )∣
4 ∣1

21
2
= +e − e ≈ 9.92.
4

Both orders of integration return the same result, as expected.

Example 13.2.2: Evaluating a double integral

Evaluate ∬
2
(3xy − x −y
2
+ 6) dA , where R is the triangle bounded by x =0 , y =0 and x/2 + y = 1 , as shown in
R

Figure 13.2.4.
Solution
While it is not specified which order we are to use, we will evaluate the double integral using both orders to help drive home
the point that it does not matter which order we use.

13.2.4 https://math.libretexts.org/@go/page/4235
Figure 13.2.4 : Finding the signed volume under the surface in Example 13.2.2 .
Using the order dy dx:
The bounds on y go from "curve to curve,'' i.e., 0 ≤ y ≤ 1 − x/2 , and the bounds on x go from "point to point,'' i.e.,
0 ≤ x ≤ 2.

x
2 − +1
2
2 2 2 2
∬ (3xy − x −y + 6) dA =∫ ∫ (3xy − x −y + 6) dy dx
R 0 0
x
2 − +1
3 1 ∣ 2
2 2 3
=∫ ( xy −x y − y + 6y) ∣ dx
0
2 3 ∣
0

2
11 11 17
3 2
=∫ ( x − x −x − ) dx
0
12 4 3

2
11 11 1 17 ∣
4 3 2
= ( x − x − x − x) ∣
48 12 2 3 ∣
0

17 ¯
¯¯
= = 5. 6.
3

Now lets consider the order dxdy . Here x goes from "curve to curve,'' 0 ≤ x ≤ 2 − 2y , and y goes from "point to point,''
0 ≤ y ≤ 1:

1 2−2y
2 2 2 2
∬ (3xy − x −y + 6) dA =∫ ∫ (3xy − x −y + 6) dx dy
R 0 0

1 2−2y
3 1 ∣
2 3 2
=∫ ( x y− x − xy + 6x) ∣ dy
0
2 3 ∣
0

1
32 3 2
28
=∫ ( y − 22 y + 2y + ) dy
0
3 3

1
8 22 28 ∣
4 3 2
= ( y − y +y + y) ∣
3 3 3 ∣ 0

17 ¯
¯¯
= = 5. 6.
3

We obtained the same result using both orders of integration.

Note how in these two examples that the bounds of integration depend only on R ; the bounds of integration have nothing to do
with f (x, y). This is an important concept, so we include it as a Key Idea.

KEY IDEA 56: Double Integration Bounds

When evaluating ∬ f (x, y)dA using an iterated integral, the bounds of integration depend only on R . The surface f does
R

not determine the bounds of integration.

Before doing another example, we give some properties of double integrals. Each should make sense if we view them in the
context of finding signed volume under a surface, over a region.

13.2.5 https://math.libretexts.org/@go/page/4235
THEOREM 120 Properties of Double Integrals
Let f and g be continuous functions over a closed, bounded plane region R , and let c be a constant.

1. ∬ c f (x, y)dA = c ∬ f (x, y) dA.


R R

2. ∬ (f (x, y) ± g(x, y)) dA = ∬ f (x, y) dA ± ∬ g(x, y) dA


R R R

3. If f (x, y) ≥ 0 on R , then ∬ f (x, y) dA ≥ 0 .


R

4. If f (x, y) ≥ g(x, y) on R , then ∬ f (x, y) dA ≥ ∬ g(x, y) dA .


R R

5. Let R be the union of two nonoverlapping regions, R = R 1 ⋃ R2 (see Figure 13.2.5). Then

∬ f (x, y) dA = ∬ f (x, y) dA + ∬ f (x, y) dA. (13.2.12)


R R1 R2

Figure 13.2.5 : R is the union of two nonoverlapping regions, R and R .1 2

Example 13.2.3: Evaluating a double integral

Let f (x, y) = sin x cos y and R be the triangle with vertices ,


(−1, 0) (1, 0) and (0, 1) (Figure 13.2.5 ). Evaluate the double
integral ∬ .
f (x, y) dA
R

Figure 13.2.5 : Finding the signed volume under a surface in Example 13.2.3 .
Solution
If we attempt to integrate using an iterated integral with the order dydx, note how there are two upper bounds on R meaning
we'll need to use two iterated integrals. We would need to split the triangle into two regions along the y -axis, then use Theorem
120, part 5.
Instead, let's use the order dxdy. The curves bounding x are y − 1 ≤ x ≤ 1 − y ; the bounds on y are 0 ≤y ≤1 . This gives
us:

13.2.6 https://math.libretexts.org/@go/page/4235
1 1−y

∬ f (x, y) dA = ∫ ∫ sin x cos y dx dy


R 0 y−1

1 1−y

=∫ ( − cos x cos y) dy

y−1
0

=∫ cos y( − cos(1 − y) + cos(y − 1)) dy.


0

Recall that the cosine function is an even function; that is, cos x = cos(−x). Therefore, from the last integral above, we have
cos(y − 1) = cos(1 − y) . Thus the integrand simplifies to 0, and we have

∬ f (x, y) dA = ∫ 0 dy
R 0

= 0.

It turns out that over R , there is just as much volume above the xy-plane as below (look again at Figure 13.2.5), giving a final
signed volume of 0.

Example 13.2.4: Evaluating a double integral

Evaluate ∬ (4 − y) dA , where R is the region bounded by the parabolas y 2


= 4x and x
2
= 4y , graphed in Figure 13.2.6.
R

Figure 13.2.6 : Finding the volume under the surface in Example 13.2.4 .
Solution
Graphing each curve can help us find their points of intersection. Solving analytically, the second equation tells us that
y = x /4 . Substituting this value in for y in the first equation gives us x /16 = 4x. Solving for x:
2 4

4
x
= 4x
16
4
x − 64x = 0
3
x(x − 64) = 0

x = 0, 4.

Thus we've found analytically what was easy to approximate graphically: the regions intersect at (0, 0) and (4, 4), as shown in
Figure 13.2.6.
We now choose an order of integration: dy dx or dx dy? Either order works; since the integrand does not contain x, choosing
dx dy might be simpler -- at least, the first integral is very simple.

Thus we have the following "curve to curve, point to point'' bounds: y 2


/4 ≤ x ≤ 2 √y , and 0 ≤ y ≤ 4 .

13.2.7 https://math.libretexts.org/@go/page/4235
4 2 √y

∬ (4 − y) dA = ∫ ∫ (4 − y) dx dy
2
R 0 y /4

4
2 √y

=∫ (x(4 − y)) dy
∣ 2
y /4
0

4 2 4 3
y y 2 3/2 1/2
=∫ ((2 √y − )(4 − y)) dy = ∫ ( −y − 2y + 8y ) dy
0
4 0
4
4
4 3 5/2 3/2
y y 4y 16y ∣
= ( − − + )∣
16 3 5 3 ∣ 0

176 ¯
¯¯
= = 11.7 3.
15

The signed volume under the surface f is about 11.7 cubic units.

In the previous section we practiced changing the order of integration of a given iterated integral, where the region R was not
explicitly given. Changing the bounds of an integral is more than just an test of understanding. Rather, there are cases where
integrating in one order is really hard, if not impossible, whereas integrating with the other order is feasible.

Example 13.2.5: Changing the order of integration


3 3
2

Rewrite the iterated integral ∫ ∫ e


−x
dx dy with the order dy dx. Comment on the feasibility to evaluate each integral.
0 y

Solution
Once again we make a sketch of the region over which we are integrating to facilitate changing the order. The bounds on x are
from x = y to x = 3 ; the bounds on y are from y = 0 to y = 3 . These curves are sketched in Figure 13.2.7, enclosing the
region R .

Figure 13.2.7 : Determining the region R determined by the bounds of integration in Example 13.2.5 .
To change the bounds, note that the curves bounding y are y =0 up to y =x ; the triangle is enclosed between x =0 and
3 x
2

x =3 . Thus the new bounds of integration are 0 ≤ y ≤ x and 0 ≤ x ≤ 3 , giving the iterated integral ∫ ∫ e
−x
dy dx .
0 0

How easy is it to evaluate each iterated integral? Consider the order of integrating dxdy, as given in the original problem. The
2

first indefinite integral we need to evaluate is ∫ e


−x
dx ; we have stated before (see Section 5.5) that this integral cannot be
evaluated in terms of elementary functions. We are stuck.
2

Changing the order of integration makes a big difference here. In the second iterated integral, we are faced with ∫ e
−x
dy ;
integrating with respect to y gives us y e , and the first definite integral evaluates to
2
−x
+C

x
2 2
−x −x
∫ e dy = x e .
0

Thus
3 x 3
2 2
−x −x
∫ ∫ e dy dx = ∫ (x e ) dx.
0 0 0

13.2.8 https://math.libretexts.org/@go/page/4235
This last integral is easy to evaluate with substitution, giving a final answer of 1

2
(1 − e
−9
) ≈ 0.5 . Figure 13.2.8 shows the
surface over R .

Figure 13.2.8 : Showing the surface f defined in Example 13.2.5 over its region R.
In short, evaluating one iterated integral is impossible; the other iterated integral is relatively simple.

Definition 22 defines the average value of a single--variable function f (x) on the interval [a, b] as
b
1
average value of f (x) on [a, b] = ∫ f (x) dx;
b −a a

that is, it is the "area under f over an interval divided by the length of the interval.'' We make an analogous statement here: the
average value of z = f (x, y) over a region R is the volume under f over R divided by the area of R .

Definition 102 The Average Value of f on R

Let z = f (x, y) be a continuous function defined over a closed region R in the xy-plane. The average value of f on R is
∬ f (x, y) dA
R
average value of f on R = . (13.2.13)
∬ dA
R

Example 13.2.6: Finding average value of a function over a region R

Find the average value of f (x, y) = 4 − y over the region R , which is bounded by the parabolas y 2
= 4x and x 2
= 4y . Note:
this is the same function and region as used in Example 13.2.4.
Solution
In Example 13.2.4 we found
4 2 √y
176
∬ f (x, y) dA = ∫ ∫ (4 − y) dx dy = .
R 0 y 2
/4 15

We find the area of R by computing ∬ dA :


R

4 2 √y
16
∬ dA = ∫ ∫ dx dy = .
R 0 y
2
/4
3

Dividing the volume under the surface by the area gives the average value:
176/15 11
average value of f on R = = = 2.2.
16/3 5

While the surface, as shown in Figure 13.2.9, covers z -values from z = 0 to z = 4 , the "average'' z -value on R is 2.2.

13.2.9 https://math.libretexts.org/@go/page/4235
Figure 13.2.9 : Finding the average value of f in Example 13.2.6 .

The previous section introduced the iterated integral in the context of finding the area of plane regions. This section has extended
our understanding of iterated integrals; now we see they can be used to find the signed volume under a surface.
This new understanding allows us to revisit what we did in the previous section. Given a region R in the plane, we computed
∬ 1 dA ; again, our understanding at the time was that we were finding the area of R . However, we can now view the function
R

z =1 as a surface, a flat surface with constant z -value of 1. The double integral ∬ 1 dA finds the volume, under z = 1 , over R ,
R

as shown in Figure 13.2.10. Basic geometry tells us that if the base of a general right cylinder has area A , its volume is A ⋅ h ,
where h is the height. In our case, the height is 1. We were "actually'' computing the volume of a solid, though we interpreted the
number as an area.

Figure 13.2.10: Showing how an iterated integral used to find area finds a certain volume.
The next section extends our abilities to find "volumes under surfaces.'' Currently, some integrals are hard to compute because
either the region R we are integrating over is hard to define with rectangular curves, or the integrand itself is hard to deal with.
Some of these problems can be solved by converting everything into polar coordinates.

This page titled 13.2: Double Integration and Volume is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by
Gregory Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available
upon request.

13.2.10 https://math.libretexts.org/@go/page/4235
13.3: Double Integration with Polar Coordinates
We have used iterated integrals to evaluate double integrals, which give the signed volume under a surface, z = f (x, y) , over a
region R of the xy-plane. The integrand is simply f (x, y), and the bounds of the integrals are determined by the region R .
Some regions R are easy to describe using rectangular coordinates -- that is, with equations of the form y = f (x), x = a , etc.
However, some regions are easier to handle if we represent their boundaries with polar equations of the form r = f (θ) , θ = α , etc.

The basic form of the double integral is ∬ . We interpret this integral as follows: over the region
f (x, y) dA R , sum up lots of
R

products of heights (given by f (x , y )) and areas (given by ΔA ). That is, dA represents "a little bit of area.'' In rectangular
i i i

coordinates, we can describe a small rectangle as having area dx dy or dy dx -- the area of a rectangle is simply length×width -- a
small change in x times a small change in y . Thus we replace dA in the double integral with dx dy or dy dx.

FIGURE 13.3.1
Now consider representing a region R with polar coordinates. Consider Figure 13.3.1a. Let R be the region in the first quadrant
bounded by the curve. We can approximate this region using the natural shape of polar coordinates: portions of sectors of circles. In
the figure, one such region is shaded, shown again in part (b) of the figure.
As the area of a sector of a circle with radius r, subtended by an angle θ , is A = r 1

2
2
θ, we can find the area of the shaded region.
The whole sector has area r Δθ, whereas the smaller, unshaded sector has area
1

2
2
2
1

2
r Δθ. The area of the shaded region is the
2
1

difference of these areas:


1 1 1 r2 + r1
2 2 2 2
ΔAi = r Δθ − r Δθ = (r − r )(Δθ) = (r2 − r1 )Δθ. (13.3.1)
2 1 2 1
2 2 2 2

Note that (r2 + r1 )/2 is just the average of the two radii.
To approximate the region R , we use many such subregions; doing so shrinks the difference r − r between radii to 0 and shrinks
2 1

the change in angle Δθ also to 0. We represent these infinitesimal changes in radius and angle as dr and dθ , respectively. Finally,
as dr is small, r ≈ r , and so (r + r )/2 ≈ r . Thus, when dr and dθ are small,
2 1 2 1 1

ΔAi ≈ ri dr dθ. (13.3.2)

Taking a limit, where the number of subregions goes to infinity and both r 2 − r1 and Δθ go to 0, we get

dA = r dr dθ. (13.3.3)

13.3.1 https://math.libretexts.org/@go/page/4236
So to evaluate ∬ , replace
f (x, y) dA dA with r dr dθ . Convert the function z = f (x, y) to a function with polar coordinates
R

with the substitutions x = r cos θ , y = r sin θ . Finally, find bounds g 1 (θ) ≤ r ≤ g2 (θ) and α ≤ θ ≤ β that describe R . This is the
key principle of this section, so we restate it here as a Key Idea.

Key Idea: Evaluating Double Integrals with Polar Coordinates


Let R be a plane region bounded by the polar equations α ≤ θ ≤ β and g 1 (θ) ≤ r ≤ g2 (θ) . Then
β g2 (θ)

∬ f (x, y) dA = ∫ ∫ f (r cos θ, r sin θ) r dr dθ. (13.3.4)


R α g (θ)
1

Examples will help us understand this Key Idea.

Example 13.3.1: Evaluating a double integral with polar coordinates

Find the signed volume under the plane z = 4 − x − 2y over the circle with equation x 2
+y
2
=1 .

Solution
The bounds of the integral are determined solely by the region R over which we are integrating. In this case, it is a circle with
equation x + y = 1 . We need to find polar bounds for this region. It may help to review polar coordinates earlier in this text;
2 2

bounds for this circle are 0 ≤ r ≤ 1 and 0 ≤ θ ≤ 2π .


We replace f (x, y) with f (r cos θ, r sin θ). That means we make the following substitutions:

4 − x − 2y ⇒ 4 − r cos θ − 2r sin θ.

Finally, we replace dA in the double integral with r dr dθ. This gives the final iterated integral, which we evaluate:
2π 1

∬ f (x, y) dA = ∫ ∫ (4 − r cos θ − 2r sin θ)r dr dθ


R 0 0

2π 1
2
=∫ ∫ (4r − r (cos θ − 2 sin θ))dr dθ
0 0

2π 1
1 ∣
2 3
=∫ (2 r − r (cos θ − 2 sin θ)) ∣ dθ
0
3 ∣
0


1
=∫ (2 − ( cos θ − 2 sin θ)) dθ
0
3


1 ∣
= (2θ − ( sin θ + 2 cos θ)) ∣
3 ∣
0

= 4π ≈ 12.566.

FIGURE 13.3.2

13.3.2 https://math.libretexts.org/@go/page/4236
The surface and region R are shown in Figure 13.3.2.

Example 13.3.2: Evaluating a double integral with polar coordinates

Find the volume under the paraboloid z = 4 − (x − 2) 2


−y
2
over the region bounded by the circles (x − 1) 2
+y
2
=1 and
(x − 2 ) + y = 4 .
2 2

Solution
At first glance, this seems like a very hard volume to compute as the region R (shown in Figure 13.3.3a) has a hole in it,
cutting out a strange portion of the surface, as shown in part (b) of the figure. However, by describing R in terms of polar
equations, the volume is not very difficult to compute.

FIGURE 13.3.3
It is straightforward to show that the circle (x − 1) + y = 1 has polar equation r = 2 cos θ , and that the circle
2 2

(x − 2 ) + y = 4 has polar equation r = 4 cos θ . Each of these circles is traced out on the interval 0 ≤ θ ≤ π . The bounds
2 2

on r are 2 cos θ ≤ r ≤ 4 cos θ.


Replacing x with r cos θ in the integrand, along with replacing y with r sin θ , prepares us to evaluate the double integral
∬ f (x, y) dA :
R

13.3.3 https://math.libretexts.org/@go/page/4236
π 4 cos θ
2 2
∬ f (x, y) dA =∫ ∫ (4 − (r cos θ − 2 ) − (r sin θ) )r dr dθ
R 0 2 cos θ

π 4 cos θ
3 2
=∫ ∫ (−r + 4r cos θ)dr dθ
0 2 cos θ

π 4 cos θ
1 4 ∣
4 3
=∫ (− r + r cos θ) ∣ dθ
0 4 3 ∣
2 cos θ
π
1 4
4 4
=∫ ([− (256 cos θ) + (64 cos θ)] −
0
4 3

1 4
4 4
[− (16 cos θ) + (8 cos θ)]) dθ
4 3
π
44
4
=∫ cos θ dθ.
0
3

To integrate cos4
θ , rewrite it as cos2
θ cos
2
θ and employ the power-reducing formula twice:
4 2 2
cos θ = cos θ cos θ

1 1
= (1 + cos(2θ)) (1 + cos(2θ))
2 2
1
2
= (1 + 2 cos(2θ) + cos (2θ))
4
1 1
= (1 + 2 cos(2θ) + (1 + cos(4θ)))
4 2
3 1 1
= + cos(2θ) + cos(4θ).
8 2 8

Picking up from where we left off above, we have


π
44
4
=∫ cos θ dθ
0
3
π
44 3 1 1
=∫ ( + cos(2θ) + cos(4θ)) dθ
0
3 8 2 8
π
44 3 1 1 ∣
= ( θ+ sin(2θ) + sin(4θ)) ∣
3 8 4 32 ∣
0

11
= π ≈ 17.279.
2

While this example was not trivial, the double integral would have been much harder to evaluate had we used rectangular
coordinates.

Example 13.3.3: Evaluating a double integral with polar coordinates


1
Find the volume under the surface f (x, y) = over the sector of the circle with radius a centered at the origin in
x2 + y 2 + 1

the first quadrant, as shown in Figure 13.3.4.

13.3.4 https://math.libretexts.org/@go/page/4236
FIGURE 13.3.4
Solution
The region R we are integrating over is a circle with radius a , restricted to the first quadrant. Thus, in polar, the bounds on R
are 0 ≤ r ≤ a , 0 ≤ θ ≤ π/2 . The integrand is rewritten in polar as
1 1 1
⇒ = . (13.3.5)
2 2 2 2 2 2 2
x +y +1 r cos θ+r sin θ+1 r +1

We find the volume as follows:


π/2 a
r
∬ f (x, y) dA = ∫ ∫ dr dθ
2
R 0 0 r +1
π/2
1 a
2 ∣
=∫ ( ln | r + 1|) dθ

0
2 0

π/2
1
2
=∫ ln(a + 1) dθ
0
2

π/2
1 ∣
2
= ( ln(a + 1)θ) ∣
2 ∣
0

π
2
= ln(a + 1).
4

Figure 13.3.4 shows that f shrinks to near 0 very quickly. Regardless, as a grows, so does the volume, without bound.

Note: Previous work has shown that there is finite area under x2 +1
1
over the entire x-axis. However, Example 13.3.3 shows
1
that there is infinite volume under 2 2
over the entire xy-plane.
x +y +1

Example 13.3.4: Finding the volume of a sphere

Find the volume of a sphere with radius a .

Solution
−−−−−−−−− −
The sphere of radius a , centered at the origin, has equation x + y + z = a ; solving for z , we have z = √a − x − y .
2 2 2 2 2 2 2

This gives the upper half of a sphere. We wish to find the volume under this top half, then double it to find the total volume.
The region we need to integrate over is the circle of radius a , centered at the origin. Polar bounds for this equation are
0 ≤ r ≤ a , 0 ≤ θ ≤ 2π .

All together, the volume of a sphere with radius a is:


−−−−−−−−−− 2π a −−−−−−−−−−−−−−−−−−−−
2 2 2 2 2 2
2∬ √a −x −y dA = 2 ∫ ∫ √a − (r cos θ) − (r sin θ) r dr dθ (13.3.6)
R 0 0

2π a
− −−−−−
2 2
=2∫ ∫ r√ a − r dr dθ.
0 0

We can evaluate this inner integral with substitution. With u =a


2
−r
2
, du = −2r dr . The new bounds of integration are
u(0) = a to u(a) = 0 . Thus we have:
2

13.3.5 https://math.libretexts.org/@go/page/4236
2π 0
1/2
=∫ ∫ (−u )du dθ
0 a2

2π 0
2 ∣
3/2
=∫ (− u )∣ dθ
0 3 ∣ 2
a


2
3
=∫ ( a ) dθ
0 3

2 ∣
3
=( a θ) ∣
3 ∣
0

4
3
= πa .
3

Generally, the formula for the volume of a sphere with radius r is given as 4/3πr ; we have justified this formula with our
3

calculation.

Example 13.3.5: Finding the volume of a solid

A sculptor wants to make a solid bronze cast of the solid shown in Figure 13.3.5, where the base of the solid has boundary, in
polar coordinates, r = cos(3θ) , and the top is defined by the plane z = 1 − x + 0.1y . Find the volume of the solid.

FIGURE 13.3.5
Solution
From the outset, we should recognize that knowing how to set up this problem is probably more important than knowing how
to compute the integrals. The iterated integral to come is not "hard'' to evaluate, though it is long, requiring lots of algebra.
Once the proper iterated integral is determined, one can use readily--available technology to help compute the final answer.
The region R that we are integrating over is bound by 0 ≤ r ≤ cos(3θ) , for 0 ≤ θ ≤ π (note that this rose curve is traced out
on the interval [0, π], not [0, 2π]). This gives us our bounds of integration. The integrand is z = 1 − x + 0.1y ; converting to
polar, we have that the volume V is:
π cos(3θ)

V =∬ f (x, y) dA = ∫ ∫ (1 − r cos θ + 0.1r sin θ)r dr dθ.


R 0 0

Distributing the r, the inner integral is easy to evaluate, leading to


π
1 1 0.1
2 3 3
∫ ( cos (3θ) − cos (3θ) cos θ + cos (3θ) sin θ) dθ.
0
2 3 3

This integral takes time to compute by hand; it is rather long and cumbersome. The powers of cosine need to be reduced, and
products like cos(3θ) cos θ need to be turned to sums using the Product To Sum formulas in the back cover of this text.
We rewrite 1

2
2
cos (3θ) as 1

4
(1 + cos(6θ)) . We can also rewrite 1

3
3
cos (3θ) cos θ as:

13.3.6 https://math.libretexts.org/@go/page/4236
1 1 1 1 + cos(6θ)
3 2
cos (3θ) cos θ = cos (3θ) cos(3θ) cos θ = ( cos(4θ) + cos(2θ)).
3 3 3 2

This last expression still needs simplification, but eventually all terms can be reduced to the form a cos(mθ) or a sin(mθ) for
various values of a and m.
We forgo the algebra and recommend the reader employ technology, such as WolframAlpha, to compute the numeric answer.
Such technology gives:
π cos(3θ)
π
3
∫ ∫ (1 − r cos θ + 0.1r sin θ)r dr dθ = ≈ 0.785 u .
0 0
4

Since the units were not specified, we leave the result as almost 0.8 cubic units (meters, feet, etc.) Should the artist want to
scale the piece uniformly, so that each rose petal had a length other than 1, she should keep in mind that scaling by a factor of k
scales the volume by a factor of k . 3

We have used iterated integrals to find areas of plane regions and volumes under surfaces. Just as a single integral can be used to
compute much more than "area under the curve,'' iterated integrals can be used to compute much more than we have thus far seen.
The next two sections show two, among many, applications of iterated integrals.

This page titled 13.3: Double Integration with Polar Coordinates is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or
curated by Gregory Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history
is available upon request.

13.3.7 https://math.libretexts.org/@go/page/4236
13.4: Center of Mass
We have used iterated integrals to find areas of plane regions and signed volumes under surfaces. A brief recap of these uses will be
useful in this section as we apply iterated integrals to compute the mass and center of mass of planar regions.

To find the area of a planar region, we evaluated the double integral ∬ dA . That is, summing up the areas of lots of little
R

subregions of R gave us the total area. Informally, we think of ∬ dA as meaning "sum up lots of little areas over R .''
R

To find the signed volume under a surface, we evaluated the double integral ∬ . Recall that the "dA " is not just a
f (x, y) dA
R

"bookend'' at the end of an integral; rather, it is multiplied by f (x, y). We regard f (x, y) as giving a height, and dA still giving an
area: f (x, y) dA gives a volume. Thus, informally, ∬ f (x, y) dA means "sum up lots of little volumes over R .''
R

We now extend these ideas to other contexts.

Mass and Weight


Consider a thin sheet of material with constant thickness and finite area. Mathematicians (and physicists and engineers) call such a
sheet a lamina. So consider a lamina, as shown in Figure 13.4.1a, with the shape of some planar region R , as shown in part (b).

Figure 13.4.1 : Illustrating the concept of a lamina.

We can write a simple double integral that represents the mass of the lamina: ∬ dm , where "dm" means "a little mass.'' That is,
R

the double integral states the total mass of the lamina can be found by "summing up lots of little masses over R .''
To evaluate this double integral, partition R into n subregions as we have done in the past. The i subregion has area ΔA .
th
i

A fundamental property of mass is that "mass=density×area.'' If the lamina has a constant density δ , then the mass of this i th

subregion is Δm = δΔA . That is, we can compute a small amount of mass by multiplying a small amount of area by the density.
i i

If density is variable, with density function δ = δ(x, y) , then we can approximate the mass of the i subregion of R by th

multiplying ΔA by δ(x , y ), where (x , y ) is a point in that subregion. That is, for a small enough subregion of R , the density
i i i i i

across that region is almost constant.


Note: Mass and weight are different measures. Since they are scalar multiples of each other, it is often easy to treat them as the
same measure. In this section we effectively treat them as the same, as our technique for finding mass is the same as for finding
weight. The density functions used will simply have different units.
The total mass M of the lamina is approximately the sum of approximate masses of subregions:
n n

M ≈ ∑ Δmi = ∑ δ(xi , yi ) ΔAi .

i=1 i=1

13.4.1 https://math.libretexts.org/@go/page/4237
Taking the limit as the size of the subregions shrinks to 0 gives us the actual mass; that is, integrating δ(x, y) over R gives the mass
of the lamina.

Definition 103 Mass of a Lamina with VarIable Density

Let δ(x, y) be a continuous density function of a lamina corresponding to a plane region R . The mass M of the lamina is

mass M = ∬ dm = ∬ δ(x, y) dA. (13.4.1)


R R

Example 13.4.1: Finding the mass of a lamina with constant density

Find the mass of a square lamina, with side length 1, with a density of δ = 3 gm/cm . 2

Solution
We represent the lamina with a square region in the plane as shown in Figure 13.4.2. As the density is constant, it does not
matter where we place the square.

Figure 13.4.2 : A region R representing a lamina in Example 13.4.1 .


Following Definition 103, the mass M of the lamina is
1 1 1 1

M =∬ 3 dA = ∫ ∫ 3 dx dy = 3 ∫ ∫ dx dy = 3 gm.
R 0 0 0 0

This is all very straightforward; note that all we really did was find the area of the lamina and multiply it by the constant
density of 3 gm/cm .2

Example 13.4.2: Finding the mass of a lamina with variable density

Find the mass of a square lamina, represented by the unit square with lower lefthand corner at the origin (see Figure 13.4.2 ),
with variable density δ(x, y) = (x + y + 2) gm/cm . 2

Solution
The variable density δ , in this example, is very uniform, giving a density of 3 in the center of the square and changing linearly.
A graph of δ(x, y) can be seen in Figure 13.4.3; notice how "same amount'' of density is above z = 3 as below. We'll comment
on the significance of this momentarily.
The mass M is found by integrating δ(x, y) over R . The order of integration is not important; we choose dx dy arbitrarily.
Thus:

13.4.2 https://math.libretexts.org/@go/page/4237
1 1

M =∬ (x + y + 2)dA =∫ ∫ (x + y + 2) dx dy
R 0 0

1 1
1 ∣
2
=∫ ( x + x(y + 2)) ∣ dy
0 2 ∣
0

1
5
=∫ ( + y) dy
0
2

1
5 1 ∣
2
= ( y+ y )∣
2 2 ∣0

= 3 gm.

Figure 13.4.3 : Graphing the density functions in Examples 13.4.1 and 13.4.2 .
It turns out that since since the density of the lamina is so uniformly distributed "above and below'' z = 3 that the mass of the
lamina is the same as if it had a constant density of 3. The density functions in Examples 13.4.1 and 13.4.2 are graphed in
Figure 13.4.3, which illustrates this concept.

Example 13.4.3: Finding the weight of a lamina with variable density

Find the weight of the lamina represented by the circle with radius 2 ft, centered at the origin, with density function
−−−−−−
δ(x, y) = (x + y + 1) lb/ft . Compare this to the weight of the same lamina with density δ(x, y) = (2 √x + y + 1) lb/ft
2 2 2 2 2

2
.

Solution
A direct application of Definition 103 states that the weight of the lamina is ∬ δ(x, y) dA . Since our lamina is in the shape
R

of a circle, it makes sense to approach the double integral using polar coordinates.
The density function δ(x, y) = x + y + 1 becomes
2 2
δ(r, θ) = (r cos θ)
2
+ (r sin θ)
2
+1 = r
2
+1 . The circle is bounded
by 0 ≤ r ≤ 2 and 0 ≤ θ ≤ 2π . Thus the weight W is:
2π 2
2
W =∫ ∫ (r + 1)r dr dθ
0 0

2π 2
1 1 ∣
4 2
=∫ ( r + r )∣ dθ
0
4 2 ∣0

=∫ (6) dθ
0

= 12π ≈ 37.70 lb.

−−−−−−
Now compare this with the density function 2
δ(x, y) = 2 √x + y
2
+1 . Converting this to polar coordinates gives
−−−−−−−−−−−−−−− −
δ(r, θ) = 2 √(r cos θ)2 + (r sin θ)2 + 1 = 2r + 1 . Thus the weight W is:

13.4.3 https://math.libretexts.org/@go/page/4237
2π 2

W =∫ ∫ (2r + 1)r dr dθ
0 0


2 1 2
3 2 ∣
=∫ ( r + r ) dθ
∣0
0 3 2

22
=∫ ( ) dθ
0
3

44
= π ≈ 46.08 lb.
3

One would expect different density functions to return different weights, as we have here. The density functions were chosen,
though, to be similar: each gives a density of 1 at the origin and a density of 5 at the outside edge of the circle, as seen in
Figure 13.4.4.

Figure 13.4.4 : Graphing the density functions in 13.4.3 . In (a) is the density function δ(x, y) = x
2
+y
2
+1 ; in (b) is
−−−−−−
2
δ(x, y) = 2√x + y
2
+1 .
−−−−−−
Notice how x 2
+y
2 2
+ 1 ≤ 2 √x + y
2
+1 over the circle; this results in less weight.

Plotting the density functions can be useful as our understanding of mass can be related to our understanding of "volume under a
surface.'' We interpreted ∬ f (x, y) dA as giving the volume under f over R ; we can understand ∬ δ(x, y) dA in the same
R R

way. The "volume'' under δ over R is actually mass; by compressing the "volume'' under δ onto the xy-plane, we get "more mass''
in some areas than others -- i.e., areas of greater density.
Knowing the mass of a lamina is one of several important measures. Another is the center of mass, which we discuss next.

Center of Mass
Consider a disk of radius 1 with uniform density. It is common knowledge that the disk will balance on a point if the point is placed
at the center of the disk. What if the disk does not have a uniform density? Through trial-and-error, we should still be able to find a
spot on the disk at which the disk will balance on a point. This balance point is referred to as the center of mass, or center of
gravity. It is though all the mass is "centered'' there. In fact, if the disk has a mass of 3 kg, the disk will behave physically as
though it were a point-mass of 3 kg located at its center of mass. For instance, the disk will naturally spin with an axis through its
center of mass (which is why it is important to "balance'' the tires of your car: if they are "out of balance'', their center of mass will
be outside of the axle and it will shake terribly).
We find the center of mass based on the principle of a weighted average. Consider a college class in which your homework
average is 90%, your test average is 73%, and your final exam grade is an 85%. Experience tells us that our final grade is not the
average of these three grades: that is, it is not:
0.9 + 0.73 + 0.85
≈ 0.837 = 83.7 (13.4.2)
3

13.4.4 https://math.libretexts.org/@go/page/4237
That is, you are probably not pulling a B in the course. Rather, your grades are weighted. Let's say the homework is worth 10% of
the grade, tests are 60% and the exam is 30%. Then your final grade is:
(0.1)(0.9) + (0.6)(0.73) + (0.3)(0.85) = 0.783 = 78.3 (13.4.3)

Each grade is multiplied by a weight.


In general, given values x 1, x2 , … , xn and weights w 1, w2 , … , wn , the weighted average of the n values is
n n

∑ wi xi / ∑ wi .

i=1 i=1

In the grading example above, the sum of the weights 0.1, 0.6 and 0.3 is 1, so we don't see the division by the sum of weights in
that instance.
How this relates to center of mass is given in the following theorem.

THEOREM 121 Center of Mass of Discrete Linear System

Let point masses m , m , … , m be distributed along the x-axis at locations x


1 2 n 1, , respectively. The center of mass
x2 , … , xn

x of the system is located at


¯¯
¯

n n

¯¯
¯
x = ∑ mi xi / ∑ mi . (13.4.4)

i=1 i=1

Example 13.4.4: Finding the center of mass of a discrete linear system

1. Point masses of 2 gm are located at x = −1 , x = 2 and x = 3 are connected by a thin rod of negligible weight. Find the
center of mass of the system.
2. Point masses of 10 gm, 2 gm and 1 gm are located at x = −1 , x = 2 and x = 3 , respectively, are connected by a thin rod
of negligible weight. Find the center of mass of the system.
Solution
1. Following Theorem 121, we compute the center of mass as:
2(−1) + 2(2) + 2(3) 4 ¯
¯¯
¯¯
¯
x = = = 1. 3. (13.4.5)
2 +2 +2 3

So the system would balance on a point placed at x = 4/3, as illustrated in Figure 13.4.5a.
2. Again following Theorem 121, we find:
10(−1) + 2(2) + 1(3) −3
¯¯
¯
x = = ≈ −0.23. (13.4.6)
10 + 2 + 1 13

Placing a large weight at the left hand side of the system moves the center of mass left, as shown in Figure 13.4.5b.

Figure 13.4.5 : Illustrating point masses along a thin rod and the center of mass.

13.4.5 https://math.libretexts.org/@go/page/4237
In a discrete system (i.e., mass is located at individual points, not along a continuum) we find the center of mass by dividing the
mass into a moment of the system. In general, a moment is a weighted measure of distance from a particular point or line. In the
case described by Theorem 121, we are finding a weighted measure of distances from the y -axis, so we refer to this as the moment
about the y -axis, represented by M . Letting M be the total mass of the system, we have x̄ = M /M .
y
¯
¯
y

We can extend the concept of the center of mass of discrete points along a line to the center of mass of discrete points in the plane
rather easily. To do so, we define some terms then give a theorem.

Definition 104 Moments about the x- and y - Axes.

Let point masses m , m


1 2, … , mn be located at points (x 1, y1 ) , (x 2, y2 ) … , (xn , yn ) , respectively, in the xy-plane.
n

1. The moment about the y -axis, M , is M y y = ∑ mi xi .

i=1
n

2. The moment about the x-axis}, M , is M x x = ∑ mi yi .

i=1

One can think that these definitions are "backwards'' as M sums up "x" distances. But remember, "x" distances are measurements
y

of distance from the y -axis, hence defining the moment about the y -axis.
We now define the center of mass of discrete points in the plane.

THEOREM 122 Center of Mass of Discrete Planar System

Let point masses m1 ,m 2, … , mn be located at points (x1 , y1 ) , (x2 , y2 ) … , (xn , yn ) , respectively, in the xy-plane, and let
n

M = ∑ mi .
i=1

The center of mass of the system is at (x̄, ¯


¯ ¯¯
ȳ ) , where
My Mx
¯¯ ¯¯
x̄ = and ȳ = . (13.4.7)
M M

Example 13.4.5: Finding the center of mass of a discrete planar system

Let point masses of 1 kg, 2 kg and 5 kg be located at points (2, 0), (1, 1) and , respectively, and are connected by thin
(3, 1)

rods of negligible weight. Find the center of mass of the system.

Solution
We follow Theorem 122 and Definition 104 to find M , M and M : x y

M = 1 +2 +5 = 8 kg.
n

Mx = ∑ mi yi

i=1

= 1(0) + 2(1) + 5(1)

= 7.

My = ∑ mi xi

i=1

= 1(2) + 2(1) + 5(3)

= 19.

13.4.6 https://math.libretexts.org/@go/page/4237
Figure 13.4.6 : Illustrating the center of mass of a discrete planar system in Example 13.4.5 .
My
Thus the center of mass is (x̄, ¯
¯ ¯¯
ȳ ) =(
M
,
Mx

M
) =(
1

9
8,
7

8
) = (2.375, 0.875), illustrated in Figure 13.4.6.

We finally arrive at our true goal of this section: finding the center of mass of a lamina with variable density. While the above
measurement of center of mass is interesting, it does not directly answer more realistic situations where we need to find the center
of mass of a contiguous region. However, understanding the discrete case allows us to approximate the center of mass of a planar
lamina; using calculus, we can refine the approximation to an exact value.
We begin by representing a planar lamina with a region R in the xy-plane with density function δ(x, y). Partition R into n
subdivisions, each with area ΔA . As done before, we can approximate the mass of the i subregion with δ(x , y ) ΔA , where
i
th
i i i

(x , y ) is a point inside the i subregion. We can approximate the moment of this subregion about the y -axis with
th
i i

x δ(x , y ) ΔA
i i i -- that is, by multiplying the approximate mass of the region by its approximate distance from the y -axis.
i

Similarly, we can approximate the moment about the x-axis with y δ(x , y ) ΔA . By summing over all subregions, we have:
i i i i

mass: M ≈ ∑ δ(xi , yi ) ΔAi (as seen before)

i=1

moment about the x-axis: Mx ≈ ∑ yi δ(xi , yi ) ΔAi

i=1

moment about the y-axis: My ≈ ∑ xi δ(xi , yi ) ΔAi

i=1

By taking limits, where size of each subregion shrinks to 0 in both the x and y directions, we arrive at the double integrals given in
the following theorem.

Theorem 123: Center of Mass of a Planar Lamina, Moments


Let a planar lamina be represented by a region R in the xy-plane with density function δ(x, y).

1. mass: M =∬ δ(x, y) dA
R

2. moment about the x-axis: M x =∬ yδ(x, y) dA


R

3. moment about the y-axis: M y =∬ xδ(x, y) dA


R

4. The center of mass of the lamina is


My Mx
¯¯
¯ ¯
¯¯
(x, y ) = ( , ).
M M

We start our practice of finding centers of mass by revisiting some of the lamina used previously in this section when finding mass.
We will just set up the integrals needed to compute M , M and M and leave the details of the integration to the reader.
x y

Example 13.4.6: Finding the center of mass of a lamina

Find the center mass of a square lamina, with side length 1, with a density of δ =3 gm/cm . (Note: this is the lamina from
2

Example 13.4.1.)

13.4.7 https://math.libretexts.org/@go/page/4237
Solution
We represent the lamina with a square region in the plane as shown in Figure 13.4.7 as done previously.

Figure 13.4.7 : A region R representing a lamina in Example 13.4.6 .


Following Theorem 123, we find M , M and M : x y

1 1

M =∬ 3 dA = ∫ ∫ 3 dx dy = 3 gm.
R 0 0

1 1

Mx =∬ 3y dA = ∫ ∫ 3y dx dy = 3/2 = 1.5.
R 0 0

1 1

My =∬ 3x dA = ∫ ∫ 3x dx dy = 3/2 = 1.5.
R 0 0

My Mx
Thus the center of mass is ¯¯
¯ ¯
¯¯
(x, y ) = (
M
,
M
) = (1.5/3, 1.5/3) = (0.5, 0.5). This is what we should have expected: the
center of mass of a square with constant density is the center of the square.

Example 13.4.7: Finding the center of mass of a lamina

Find the center of mass of a square lamina, represented by the unit square with lower left-hand corner at the origin (see Figure
13.4.7), with variable density δ(x, y) = (x + y + 2) gm/cm . (Note: this is the lamina from Example 13.4.2.)
2

Solution
We follow Theorem 123, to find M , M and M : x y

1 1

M =∬ (x + y + 2) dA = ∫ ∫ (x + y + 2) dx dy = 3 gm.
R 0 0

1 1
19
Mx =∬ y(x + y + 2) dA = ∫ ∫ y(x + y + 2) dx dy = .
R 0 0
12
1 1
19
My =∬ x(x + y + 2) dA = ∫ ∫ x(x + y + 2) dx dy = .
R 0 0 12

My Mx
Thus the center of mass is (x̄, ȳ) = (
¯
¯ ¯
¯
M
,
M
) =(
19

36
,
19

36
) ≈ (0.528, 0.528). While the mass of this lamina is the same as the
lamina in the previous example, the greater density found with greater x and y values pulls the center of mass from the center
slightly towards the upper right-hand corner.

Example 13.4.8: Finding the center of mass of a lamina

Find the center of mass of the lamina represented by the circle with radius 2 ft, centered at the origin, with density function
δ(x, y) = (x + y + 1) lb/ft . (Note: this is one of the lamina used in Example 13.4.3.)
2 2 2

Solution
As done in Example 13.4.3, it is best to describe R using polar coordinates.
Thus when we compute M , we will integrate not xδ(x, y) = x(x
y
2
+y
2
+ 1) , but rather
(r cos θ)δ(r cos θ, r sin θ) = (r cos θ)(r + 1). We compute M , M and M :
2
x y

13.4.8 https://math.libretexts.org/@go/page/4237
2π 2
2
M =∫ ∫ (r + 1)r dr dθ = 12π ≈ 37.7 lb.
0 0

2π 2
2
Mx = ∫ ∫ (r sin θ)(r + 1)r dr dθ = 0.
0 0

2π 2
2
My =∫ ∫ (r cos θ)(r + 1)r dr dθ = 0.
0 0

Since R and the density of R are both symmetric about the x and y axes, it should come as no big surprise that the moments
about each axis is 0. Thus the center of mass is (x̄, ȳ) = (0, 0) .
¯
¯ ¯
¯

Example 13.4.9: Finding the center of mass of a lamina

Find the center of mass of the lamina represented by the region R shown in Figure 13.4.8, half an annulus with outer radius 6
and inner radius 5, with constant density 2 lb/ft . 2

Solution
Once again it will be useful to represent R in polar coordinates. Using the description of R and/or the illustration, we see that
R is bounded by 5 ≤ r ≤ 6 and 0 ≤ θ ≤ π . As the lamina is symmetric about the y -axis, we should expect M = 0 . We y

compute M , M and M :
x y

π 6

M =∫ ∫ (2)r dr dθ = 11π lb.


0 5

π 6
364
Mx = ∫ ∫ (r sin θ)(2)r dr dθ = ≈ 121.33.
0 5
3
π 6

My = ∫ ∫ (r cos θ)(2)r dr dθ = 0.
0 5

Figure 13.4.8: Illustrating the region R in Example 13.4.9.


Thus the center of mass is ¯¯
(x̄ ¯¯
, ȳ ) = (0,
364

33π
) ≈ (0, 3.51). The center of mass is indicated in Figure ; note how it lies
13.4.8

outside of R !

This section has shown us another use for iterated integrals beyond finding area or signed volume under the curve. While there are
many uses for iterated integrals, we give one more application in the following section: computing surface area.

This page titled 13.4: Center of Mass is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et
al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

13.4.9 https://math.libretexts.org/@go/page/4237
13.5: Surface Area
In Section 7.4 we used definite integrals to compute the arc length of plane curves of the form y = f (x). We later extended these
ideas to compute the arc length of plane curves defined by parametric or polar equations.
The natural extension of the concept of "arc length over an interval'' to surfaces is "surface area over a region.''
Consider the surface z = f (x, y) over a region R in the xy-plane, shown in Figure 13.5.1a. Because of the domed shape of the
surface, the surface area will be greater than that of the area of the region R . We can find this area using the same basic technique
we have used over and over: we'll make an approximation, then using limits, we'll refine the approximation to the exact value.

Figure 13.5.1 : Developing a method of computing surface area.


As done to find the volume under a surface or the mass of a lamina, we subdivide R into n subregions. Here we subdivide R into
rectangles, as shown in the figure. One such subregion is outlined in the figure, where the rectangle has dimensions Δx and Δy , i i

along with its corresponding region on the surface.


In part (b) of the figure, we zoom in on this portion of the surface. When Δx and Δy are small, the function is approximated well
i i

by the tangent plane at any point (x , y ) in this subregion, which is graphed in part (b). In fact, the tangent plane approximates the
i i

function so well that in this figure, it is virtually indistinguishable from the surface itself! Therefore we can approximate the surface
area S of this region of the surface with the area T of the corresponding portion of the tangent plane.
i i

This portion of the tangent plane is a parallelogram, defined by sides u⃗ and v ⃗, as shown. One of the applications of the cross
product from Section 10.4 is that the area of this parallelogram is ∥u⃗ × v∥⃗ . Once we can determine u⃗ and v ⃗, we can determine the
area.
u⃗is tangent to the surface in the direction of x, therefore, from Section 12.7, u⃗ is parallel to ⟨1, 0, f (x , y )⟩. The x-displacement
x i i

of u⃗ is Δx , so we know that u⃗ = Δx ⟨1, 0, f (x , y )⟩ . Similar logic shows that v ⃗ = Δy ⟨0, 1, f (x , y )⟩ . Thus:


i i x i i i y i i

surface area Si ≈ area of Ti

= ∥ u⃗ × v∥

=∣∣Δxi ⟨1, 0, fx (xi , yi )⟩ × Δyi ⟨0, 1, fy (xi , yi )⟩∣


∣∣ ∣∣

−−−−−−−−−−−−−−−−−−−−−
2 2
= √ 1 + fx (xi , yi ) + fy (xi , yi ) Δxi Δyi .

Note that Δx i Δyi = ΔAi , the area of the i th


subregion.
Summing up all n of the approximations to the surface area gives
n
−−−−−−−−−−−−−−−−−−−−−
2 2
surface area over R ≈ ∑ √ 1 + fx (xi , yi ) + fy (xi , yi ) ΔAi . (13.5.1)

i=1

Once again take a limit as all of the Δx and Δy shrink to 0; this leads to a double integral.
i i

13.5.1 https://math.libretexts.org/@go/page/4238
Definition 105 Surface Area
Let z = f (x, y) where f and f are continuous over a closed, bounded region R . The surface area S over R is
x y

S =∬ dS
R

−−−−−−−−−−−−−−−−−−−
2 2
=∬ √ 1 + fx (x, y ) + fy (x, y ) dA.
R

Note: as done before, we think of "∬ dS '' as meaning "sum up lots of little surface areas over R .''
R

The concept of surface area is defined here, for while we already have a notion of the area of a region in the plane, we did not yet
have a solid grasp of what "the area of a surface in space'' means.
We test this definition by using it to compute surface areas of known surfaces. We start with a triangle.

Example 13.5.1: Finding the surface area of a plane over a triangle

Let f (x, y) = 4 − x − 2y , and let R be the region in the plane bounded by x = 0 , y = 0 and y = 2 − x/2 , as shown in Figure
13.5.2. Find the surface area of f over R .

Figure 13.5.2 : Finding the area of a triangle in space in Example 13.5.1 .


Solution
We follow Definition 105. We start by noting that fx (x, y) = −1 and fy (x, y) = −2 . To define R , we use bounds
0 ≤ y ≤ 2 − x/2 and 0 ≤ x ≤ 4 . Therefore

S =∬ dS
R

4 2−x/2 −−−−−−−−−−−−−−−
2 2
=∫ ∫ √ 1 + (−1 ) + (−2 ) dy dx
0 0

4
– x
=∫ √6 (2 − ) dx
0
2

= 4 √6.

Because the surface is a triangle, we can figure out the area using geometry. Considering the base of the triangle to be the side
− −
in the xy-plane, we find the length of the base to be √20 . We can find the height using our knowledge of vectors: let u⃗ be the
−−−
side in the x-z plane and let v ⃗ be the side in the xy-plane. The height is then ∥u⃗ − proj u⃗∥ = 4√6/5 . Geometry states that
v⃗

the area is thus


1 −−−
−− –
⋅ 4 √6/5 ⋅ √20 = 4 √6. (13.5.2)
2

We affirm the validity of our formula.

It is "common knowledge'' that the surface area of a sphere of radius r is 4πr . We confirm this in the following example, which
2

involves using our formula with polar coordinates.

13.5.2 https://math.libretexts.org/@go/page/4238
Example 13.5.2: The surface area of a sphere.

Find the surface area of the sphere with radius a centered at the origin, whose top hemisphere has equation
−−−−−−−−− −
2
f (x, y) = √a − x − y
2
. 2

Solution
We start by computing partial derivatives and find
−x −y
fx (x, y) = and fy (x, y) = . (13.5.3)
− −−−−−−−− − − −−−−−−−− −
√ a2 − x2 − y 2 √ a2 − x2 − y 2

As our function f only defines the top upper hemisphere of the sphere, we double our surface area result to get the total area:
−−−−−−−−−−−−−−−−−−−
2 2
S =2∬ √ 1 + fx (x, y ) + fy (x, y ) dA
R
−−−−−−−−−−−−−−
2 2
x +y
=2∬ √1 + dA.
2 2 2
R a −x −y

The region R that we are integrating over is the circle, centered at the origin, with radius a : x + y = a . Because of this 2 2 2

region, we are likely to have greater success with our integration by converting to polar coordinates. Using the substitutions
x = r cos θ , y = r sin θ , dA = r dr dθ and bounds 0 ≤ θ ≤ 2π and 0 ≤ r ≤ a , we have:

−−−−−−−−−−−−−−−−−−−−−−
2π a 2 2 2 2
r cos θ+r sin θ
S =2∫ ∫ √1 + r dr dθ
2
0 0 a2 − r2 cos2 θ − r2 sin θ
−−−−−−−−−−
2π a
r2
=2∫ ∫ r√ 1 + dr dθ
2 2
0 0 a −r

−−−−−−−
2π a 2
a
=2∫ ∫ r√ dr dθ. (13.5.4)
2
0 0
a − r2

Apply substitution u = a 2 2
−r and integrate the inner integral, giving

2
=2∫ a dθ
(13.5.5)
0

2
= 4π a .

Our work confirms our previous formula.

−−−−−−−
a 2
a
Note: The inner integral in Equation 13.5.4 is an improper integral, as the integrand of ∫ r√
2 2
dr is not defined at
0 a −r

r =a . To properly evaluate this integral, one must use the techniques of Section 6.8.
−−−−−−−−−−
The reason this need arises is that the function f (x, y) = √a 2 2
−x −y
2
fails the requirements of Definition 105, as fx and fy

are not continuous on the boundary of the circle x + y = a . 2 2 2

The computation of the surface area is still valid. The definition makes stronger requirements than necessary in part to avoid the
use of improper integration, as when f and/or f are not continuous, the resulting improper integral may not converge. Since the
x y

improper integral does converge in this example, the surface area is accurately computed.

Example 13.5.3: Finding the surface area of a cone

h −−−−− −
The general formula for a right cone with height h and base radius a is f (x, y) = h − √x2 + y 2 , shown in Figure 13.34.
a
Find the surface area of this cone.

13.5.3 https://math.libretexts.org/@go/page/4238
Figure 13.5.3 : Finding the surface area of a cone in Example 13.5.3 .
Solution
We begin by computing partial derivatives.
xh yh
fx (x, y) = − and − .
− −−−−− − −−−−−
2 2 2 2
a√ x + y a√ x + y

Since we are integrating over the circle x


2
+y
2
=a
2
, we again use polar coordinates. Using the standard substitutions, our
integrand becomes
−−−−−−−−−−−−−−−−−−−−−−−−−
2 2
hr cos θ hr sin θ
√1 +( ) +(

− − )
− .
a√r2 a√r2

This may look intimidating at first, but there are lots of simple simplifications to be done. It amazingly reduces to just
−−−−−−
2
h 1 −−−−−−
2 2
√1 + = √a + h .
a2 a

Our polar bounds are 0 ≤ θ ≤ 2π and 0 ≤ r ≤ a . Thus


2π a
1 −−−−−−
2 2
S =∫ ∫ r √a + h dr dθ
0 0 a
2π a
1 1 − −−−−− ∣
2 2 2
=∫ ( r √ a + h )∣ dθ
0
2 a ∣0


1 − −−−−−
2 2
=∫ a√ a + h dθ
0
2
− −−−−−
2 2
= πa√ a + h .

This matches the formula found in the back of this text.

Note: Note that once again f and f are not continuous on the domain of f , as both are undefined at (0, 0). (A similar problem
x y

occurred in the previous example.) Once again the resulting improper integral converges and the computation of the surface area is
valid.

Example 13.5.4: Finding surface area over a region

Find the area of the surface f (x, y) = x


2
− 3y + 3 over the region R bounded by −x ≤ y ≤ x , 0 ≤x ≤4 , as pictured in
Figure 13.5.4.

13.5.4 https://math.libretexts.org/@go/page/4238
Figure 13.5.4 : Graphing the surface in Example 13.5.4 .
Solution
It is straightforward to compute f x (x, y) = 2x and f y (x, y) = −3 . Thus the surface area is described by the double integral
−−−−−−−−−−−−−− − −−−− −−
2 2 2
∬ √ 1 + (2x ) + (−3 ) dA = ∬ √ 10 + 4x dA.

R R

As with integrals describing arc length, double integrals describing surface area are in general hard to evaluate directly because
of the square-root. This particular integral can be easily evaluated, though, with judicious choice of our order of integration.
−−−−−−−
Integrating with order dx dy requires us to evaluate ∫ √10 + 4x dx . This can be done, though it involves Integration By
2

−−−−− −−
Parts and sinh x . Integrating with order dy dx has as its first integral ∫ √10 + 4x dy , which is easy to evaluate: it is
−1 2

−−− −−−−
simply y √10 + 4x + C . So we proceed with the order dy dx; the bounds are already given in the statement of the problem.
2

4 x
− −−−− −− − −−−− −−
2 2
∬ √ 10 + 4x dA = ∫ ∫ √ 10 + 4x dy dx

R 0 −x

4
− −−−− −− x
2
=∫ (y √ 10 + 4x )∣ dx
∣−x
0

4
− −−−− −−
2
=∫ (2x √ 10 + 4x ) dx.
0

Apply substitution with u = 10 + 4x : 2

4
1 3/2 ∣
2
= ( (10 + 4 x ) )∣
6 ∣
0

1 −− −− 2
= (37 √74 − 5 √10) ≈ 100.825 units .
3

So while the region R over which we integrate has an area of 16 units , the surface has a much greater area as its z -values
2

change dramatically over R .

In practice, technology helps greatly in the evaluation of such integrals. High powered computer algebra systems can compute
integrals that are difficult, or at least time consuming, by hand, and can at the least produce very accurate approximations with
numerical methods. In general, just knowing how to set up the proper integrals brings one very close to being able to compute the
needed value. Most of the work is actually done in just describing the region R in terms of polar or rectangular coordinates. Once
this is done, technology can usually provide a good answer.
We have learned how to integrate integrals; that is, we have learned to evaluate double integrals. In the next section, we learn how
to integrate double integrals -- that is, we learn to evaluate triple integrals, along with learning some uses for this operation.

This page titled 13.5: Surface Area is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et al.
via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

13.5.5 https://math.libretexts.org/@go/page/4238
13.6: Volume Between Surfaces and Triple Integration
We learned in Section 13.2 how to compute the signed volume V under a surface z = f (x, y) over a region R :
V =∬ f (x, y)dA . It follows naturally that if f (x, y) ≥ g(x, y) on R , then the volume between f (x, y) and g(x, y) on R is
R

V =∬ f (x, y)dA − ∬ g(x, y)dA (13.6.1)


R R

=∬ (f (x, y) − g(x, y))dA. (13.6.2)


R

theorem 124: Volume Between Surfaces


Let f and g be continuous functions on a closed, bounded region R , where f (x, y) ≥ g(x, y) for all (x, y) in R . The volume
V between f and g over R is

V =∬ (f (x, y) − g(x, y))dA. (13.6.3)


R

Example 13.6.1: Finding volume between surfaces

Find the volume of the space region bounded by the planes z = 3x + y − 4 and z = 8 − 3x − 2y in the 1
st
octant. In Figure
13.36(a) the planes are drawn; in (b), only the defined region is given.
Solution
We need to determine the region R over which we will integrate. To do so, we need to determine where the planes intersect.
They have common z -values when 3x + y − 4 = 8 − 3x − 2y . Applying a little algebra, we have:
3x + y − 4 = 8 − 3x − 2y

6x + 3y = 12

2x + y = 4

The planes intersect along the line 2x + y = 4 . Therefore the region R is bounded by x = 0 , y = 0 , and y = 4 − 2x ; we can
convert these bounds to integration bounds of 0 ≤ x ≤ 2 , 0 ≤ y ≤ 4 − 2x . Thus

V =∬ (8 − 3x − 2y − (3x + y − 4))dA
R

2 4−2x

=∫ ∫ (12 − 6x − 3y)dy dx
0 0

3
= 16 u .

The volume between the surfaces is 16 cubic units.

Figure 13.36: Finding the volume between the planes given in Example 13.6.1.

In the preceding example, we found the volume by evaluating the integral


2 4−2x

∫ ∫ (8 − 3x − 2y − (3x + y − 4))dy dx. (13.6.4)


0 0

13.6.1 https://math.libretexts.org/@go/page/4239
Note how we can rewrite the integrand as an integral, much as we did in Section 13.1:
8−3x−2y

8 − 3x − 2y − (3x + y − 4) = ∫ dz. (13.6.5)


3x+y−4

Thus we can rewrite the double integral that finds volume as


2 4−2x 2 4−2x 8−3x−2y

∫ ∫ (8 − 3x − 2y − (3x + y − 4))dy dx = ∫ ∫ (∫ dz) dy dx. (13.6.6)


0 0 0 0 3x+y−4

This no longer looks like a "double integral,'' but more like a "triple integral.'' Just as our first introduction to double integrals was
in the context of finding the area of a plane region, our introduction into triple integrals will be in the context of finding the volume
of a space region.

Figure 13.37: Approximating the volume of a region D in space.


To formally find the volume of a closed, bounded region D in space, such as the one shown in Figure 13.37(a), we start with an
approximation. Break D into n rectangular solids; the solids near the boundary of D may possibly not include portions of D and/or
include extra space. In Figure 13.37(b), we zoom in on a portion of the boundary of D to show a rectangular solid that contains
space not in D; as this is an approximation of the volume, this is acceptable and this error will be reduced as we shrink the size of
our solids.
The volume ΔV of the i solid D is ΔV = Δx Δy Δz , where Δx , Δy and Δz give the dimensions of the rectangular solid
i
th
i i i i i i i i

in the x, y and z directions, respectively. By summing up the volumes of all n solids, we get an approximation of the volume V of
D:

$$V \approx \sum_{i=1}^n \Delta V_i = \sum_{i=1}^n \Delta x_i\Delta y_i\Delta z_i.\]
Let ||ΔD|| represent the length of the longest diagonal of rectangular solids in the subdivision of D. As ||ΔD|| → 0, the volume
of each solid goes to 0, as do each of Δx , Δy and Δz , for all i. Our calculus experience tells us that taking a limit as
i i i

||ΔD|| → 0 turns our approximation of V into an exact calculation of V . Before we state this result in a theorem, we use a

definition to define some terms.

Definition 106: Triple Integrals, Iterated Integration (Part I)


Let D be a closed, bounded region in space. Let a and b be real numbers, let g1 (x) and g2 (x) be continuous functions of x,
and let f (x, y) and f (x, y) be continuous functions of x and y .
1 2

1. The volume V of D is denoted by a triple integral,

13.6.2 https://math.libretexts.org/@go/page/4239
V =∭ dV . (13.6.7)
D

b g (x) f (x,y)
2. The iterated integral ∫
a

2

g1 (x)

2

f1 (x,y)
dz dy dx is evaluated as

b g2 (x) f2 (x,y) b g2 (x) f2 (x,y)

∫ ∫ ∫ dz dy dx = ∫ ∫ (∫ dz) dy dx. (13.6.8)


a g1 (x) f1 (x,y) a g1 (x) f1 (x,y)

Evaluating the above iterated integral is triple integration.

Our informal understanding of the notation ∭ D


dV is "sum up lots of little volumes over D,'' analogous to our understanding of
∬ dA and ∬ dm .
R R

We now state the major theorem of this section.

theorem 125 Triple Integration (Part I)


Let D be a closed, bounded region in space and let ΔD be any subdivision of D into n rectangular solids, where the i
th

subregion D has dimensions Δx × Δy × Δz and volume ΔV .


i i i i i

1. The volume V of D is
n n

V =∭ dV = lim ∑ ΔVi = lim ∑ Δxi Δyi Δzi . (13.6.9)


D ||ΔD||→0 ||ΔD||→0
i=1 i=1

2. If D is defined as the region bounded by the planes x = a and x = b , the cylinders y = g x) and y = g (x) , and the ( 2

surfaces z = f (x, y) and z = f (x, y) , where a < b , g (x) ≤ g (x) and f (x, y) ≤ f (x, y) on D, then
1 2 1 2 1 2

b g2 (x) f2 (x,y)

∭ dV = ∫ ∫ ∫ dz dy dx. (13.6.10)
D a g (x) f (x,y)
1 1

3. V can be determined using iterated integration with other orders of integration (there are 6 total), as long as D is defined by
the region enclosed by a pair of planes, a pair of cylinders, and a pair of surfaces.

We evaluated the area of a plane region R by iterated integration, where the bounds were "from curve to curve, then from point to
point.'' Theorem 125 allows us to find the volume of a space region with an iterated integral with bounds "from surface to surface,
then from curve to curve, then from point to point.'' In the iterated integral
b g2 (x) f2 (x,y)

∫ ∫ ∫ dz dy dx, (13.6.11)
a g1 (x) f1 (x,y)

the bounds a ≤ x ≤ b and g (x) ≤ y ≤ g (x) define a region R in the x-y plane over which the region D exists in space.
1 2

However, these bounds are also defining surfaces in space; x = a is a plane and y = g (x) is a cylinder. The combination of these 1

6 surfaces enclose, and define, D.


Examples will help us understand triple integration, including integrating with various orders of integration.

Example 13.6.2: Finding the volume of a space region with triple integration

Find the volume of the space region in the 1 octant bounded by the plane z = 2 − y/3 − 2x/3 , shown in Figure 13.38(a),
st

using the order of integration dz dy dx . Set up the triple integrals that give the volume in the other 5 orders of integration.
Solution
Starting with the order of integration dz dy dx , we need to first find bounds on z . The region D is bounded below by the plane
z = 0 (because we are restricted to the first octant) and above by z = 2 − y/3 − 2x/3 ; 0 ≤ z ≤ 2 − y/3 − 2x/3 .

To find the bounds on y and x, we "collapse'' the region onto the x-y plane, giving the triangle shown in Figure 13.38(b). (We
know the equation of the line y = 6 − 2x in two ways. First, by setting z = 0 , we have 0 = 2 − y/3 − 2x/3 ⇒ y = 6 − 2x .
Secondly, we know this is going to be a straight line between the points (3, 0) and (0, 6) in the x-y plane.)

13.6.3 https://math.libretexts.org/@go/page/4239
Figure 13.38: The region D used in Example 13.6.2 in (a); in (b), the region found by collapsing D onto the x-y plane.
We define that region R , in the integration order of dy dx , with bounds 0 ≤ y ≤ 6 − 2x and 0 ≤ x ≤ 3 . Thus the volume V

of the region D is:

V =∭ dV
D
1 2
3 6−2x 2− y− x
3 3

=∫ ∫ ∫ dzdydz
0 0 0
1 2
3 6−2x 2− y− x
3 3

=∫ ∫ (∫ dz) dydz
0 0 0

3 6−2x 1 2
2− y− x
∣ 3 3
=∫ ∫ z dydz

0
0 0

3 6−2x
1 2
=∫ ∫ (2 − y− x) dydz.
0 0
3 3

From this step on, we are evaluating a double integral as done many times before. We skip these steps and give the final
volume,
3
= 6u . (13.6.12)

The order dz dx dy :
Now consider the volume using the order of integration dz dx dy . The bounds on z are the same as before,
0 ≤ z ≤ 2 − y/3 − 2x/3 . Collapsing the space region on the x-y plane as shown in Figure 13.38(b), we now describe this

triangle with the order of integration dxdy. This gives bounds 0 ≤ x ≤ 3 − y/2 and 0 ≤ y ≤ 6 . Thus the volume is given by
the triple integral
1 1 2
6 3− y 2− y− x
2 3 3

V =∫ ∫ ∫ dz dx dy. (13.6.13)
0 0 0

The order dx dy dz:


Following our "surface to surface…'' strategy, we need to determine the x-surfaces that bound our space region. To do so,
approach the region "from behind,'' in the direction of increasing x. The first surface we hit as we enter the region is the y -z
plane, defined by x = 0 . We come out of the region at the plane z = 2 − y/3 − 2x/3 ; solving for x, we have
x = 3 − y/2 − 3z/2 . Thus the bounds on x are: 0 ≤ x ≤ 3 − y/2 − 3z/2 .

13.6.4 https://math.libretexts.org/@go/page/4239
Figure 13.39: The region D in Example 13.6.2 is collapsed onto the y-z plane in (a); in (b), the region is collapsed onto the x-z
plane.
Now collapse the space region onto the y -z plane, as shown in Figure 13.39(a). (Again, we find the equation of the line
z = 2 − y/3 by setting x = 0 in the equation x = 3 − y/2 − 3z/2 .) We need to find bounds on this region with the order

dydz . The curves that bound y are y = 0 and y = 6 − 3z ; the points that bound z are 0 and 2. Thus the triple integral giving

volume is:

0 ≤ x ≤ 3 − y/2 − 3z/2
2 6−3z 3−y/2−3z/2
0 ≤ y ≤ 6 − 3z ⇒ ∫ ∫ ∫ dx dy dz. (13.6.14)
0 0 0

0 ≤z ≤2

The order dx dz dy :
The x-bounds are the same as the order above. We now consider the triangle in Figure 13.39(a) and describe it with the order
dzdy : 0 ≤ z ≤ 2 − y/3 and 0 ≤ y ≤ 6 . Thus the volume is given by:

$$
\begin{array}{cc}
\begin{array}{c}
0\leq x\leq 3-y/2-3z/2\\
0\leq z\leq 2-y/3\\
0\leq y\leq 6
\end{array}

&
\Rightarrow \quad \int_0^6\int_0^{2-y/3}\int_0^{3-y/2-3z/2} dx \, dz \, dy.
\end{array}
\]
The order dydzdx :
We now need to determine the y -surfaces that determine our region. Approaching the space region from "behind'' and moving
in the direction of increasing y , we first enter the region at y = 0 , and exit along the plane z = 2 − y/3 − 2x/3 . Solving for y ,
this plane has equation y = 6 − 2x − 3z . Thus y has bounds 0 ≤ y ≤ 6 − 2x − 3z .
Now collapse the region onto the x-z plane, as shown in Figure 13.39(b). The curves bounding this triangle are z =0 and
z = 2 − 2x/3 ; x is bounded by the points x = 0 to x = 3 . Thus the triple integral giving volume is:

$$

\begin{array}{cc}
\begin{array}{c}
0\leq y\leq 6-2x-3z\\
0\leq z\leq 2-2x/3\\
0\leq x\leq 3
\end{array}

&
\Rightarrow \quad \int_0^3\int_0^{2-2x/3}\int_0^{6-2x-3z} dy dz dx.

13.6.5 https://math.libretexts.org/@go/page/4239
\end{array}
\]
The order dy dx dz:
The y -bounds are the same as in the order above. We now determine the bounds of the triangle in Figure 13.39(b) using the
order dy dx dz. x is bounded by x = 0 and x = 3 − 3z/2 ; z is bounded between z = 0 and z = 2 . This leads to the triple
integral:
0 ≤ y ≤ 6 − 2x − 3z
2 3−3z/2 6−2x−3z
0 ≤ x ≤ 3 − 3z/2 ⇒ ∫ ∫ ∫ dy dx dz. (13.6.15)
0 0 0

0 ≤z ≤2

This problem was long, but hopefully useful, demonstrating how to determine bounds with every order of integration to
describe the region D. In practice, we only need 1, but being able to do them all gives us flexibility to choose the order that
suits us best.

In the previous example, we collapsed the surface into the x-y , x-z , and y -z planes as we determined the "curve to curve, point to
point'' bounds of integration. Since the surface was a triangular portion of a plane, this collapsing, or projecting, was simple: the
projection of a straight line in space onto a coordinate plane is a line.
The following example shows us how to do this when dealing with more complicated surfaces and curves.

Example 13.6.3: Finding the projection of a curve in space onto the coordinate planes

Consider the surfaces z = 3 − x − y and z = 2y , as shown in Figure 13.40(a). The curve of their intersection is shown,
2 2

along with the projection of this curve into the coordinate planes, shown dashed. Find the equations of the projections into the
coordinate planes.

Figure 13.40: Finding the projections of the curve of intersection in Example 13.6.3.
Solution
The two surfaces are z = 3 − x − y and z = 2y . To find where they intersect, it is natural to set them equal to each other:
2 2

3 − x − y = 2y . This is an implicit function of x and y that gives all points (x, y) in the x-y plane where the z values of the
2 2

two surfaces are equal.


We can rewrite this implicit function by completing the square:
2 2 2 2 2 2
3 −x −y = 2y ⇒ y + 2y + x =3 ⇒ (y + 1 ) +x = 4. (13.6.16)

Thus in the x-y plane the projection of the intersection is a circle with radius 2, centered at (0, −1).
To project onto the x-z plane, we do a similar procedure: find the x and z values where the y values on the surface are the
same. We start by solving the equation of each surface for y . In this particular case, it works well to actually solve for y : 2

2 2 2 2
z = 3 −x −y ⇒ y = 3 −x −z

z = 2y ⇒ y
2 2
= z /4 .
Thus we have (after again completing the square):

13.6.6 https://math.libretexts.org/@go/page/4239
2 2
(z + 2) x
2 2
3 −x − z = z /4 ⇒ + = 1, (13.6.17)
16 4

and ellipse centered at (0, −2) in the x-z plane with a major axis of length 8 and a minor axis of length 4.
Finally, to project the curve of intersection into the y -z plane, we solve equation for x. Since z = 2y is a cylinder that lacks the
variable x, it becomes our equation of the projection in the y -z plane.
All three projections are shown in Figure 13.40(b).

Example 13.6.4: Finding the volume of a space region with triple integration

Set up the triple integrals that find the volume of the space region D bounded by the surfaces x 2
+y
2
=1 , z =0 and
z = −y , as shown in Figure 13.41(a), with the orders of integration dz dy dx , dy dx dz and dx dz dy .

Figure 13.41: The region D in Example 13.6.4 is shown in (a); in (b), it is collapsed onto the x-y plane.
Solution
The order dz dy dx :
The region D is bounded below by the plane z = 0 and above by the plane z = −y . The cylinder x 2
+y
2
=1 does not offer
any bounds in the z -direction, as that surface is parallel to the z -axis. Thus 0 ≤ z ≤ −y .
Collapsing the region into the x-y plane, we get part of the circle with equation x + y = 1 as shown in Figure 13.41(b). As a
2 2

−−−− − −−−−−
function of x, this half circle has equation y = −√1 − x . Thus y is bounded below by −√1 − x and above by y = 0 :
2 2

−−−− −
≤ y ≤ 0 . The x bounds of the half circle are −1 ≤ x ≤ 1 . All together, the bounds of integration and triple
2
−√1 − x

integral are as follows:


0 ≤ z ≤ −y
− −−− − 1 0 −y
−√ 1 − x2 ≤ y ≤ 0 ⇒ ∫
−1

−√1−x2

0
dz dy dx. (13.6.18)

−1 ≤ x ≤ 1

We evaluate this triple integral:


1 0 −y 1 0

∫ ∫ ∫ dz dy dx = ∫ ∫ ( − y)dy dx
2 2
−1 −√1−x 0 −1 −√1−x

1
1 0
2 ∣
=∫ (− y ) dx
∣ 2
2 −√1−x
−1

1
1 2
=∫ (1 − x )dx
−1
2

1
1 1 ∣
3
= ( (x − x )) ∣
2 3 ∣−1

2
3
= units .
3

With the order dy dx dz:

13.6.7 https://math.libretexts.org/@go/page/4239
−−−− −
The region is bounded "below'' in the y -direction by the surface x 2
+y
2
= 1 ⇒ y = −√1 − x
2
and "above'' by the surface
−−−− −
y = −z . Thus the y bounds are −√1 − x ≤ y ≤ −z .
2

Figure 13.42: The region D in Example 13.6.4 is shown collapsed onto the x-z plane in (a); in (b), it is collapsed onto the y-z
plane.
Collapsing the region onto the x-z plane gives the region shown in Figure 13.42(a); this half circle has equation x + z = 1 . 2 2

(We find this curve by solving each surface for y , then setting them equal to each other. We have y = 1 − x and
2 2 2

− −−−− −− − −−
y = −z ⇒ y = z
2
. Thus x + z = 1 .) It is bounded below by x = −√1 − z and above by x = √1 − z , where z is
2 2 2 2 2

bounded by 0 ≤ z ≤ 1 . All together, we have:


− −−− −
2
−√ 1 − x ≤ y ≤ −z
− −−− − −−−− − 1 √1−z 2 −z
2 2
−√ 1 − z ≤ x ≤ √1 − z ⇒ ∫
0
∫ 2

−√1−x2
dy dx dz. (13.6.19)
−√1−z

0 ≤z ≤1

With the order dx dz dy :


−−−−− −−−−−
D is bounded below by the surface x = −√1 − y and above by √1 − y . We then collapse the region onto the y -z plane
2 2

and get the triangle shown in Figure 13.42}(b). (The hypotenuse is the line z = −y , just as the plane.) Thus z is bounded by
0 ≤ z ≤ −y and y is bounded by −1 ≤ y ≤ 0 . This gives:

− −−− − −−−− −
2 2
−√ 1 − y ≤ x ≤ √1 − y
0 −y √1−y 2

0 ≤ z ≤ −y ⇒ ∫ ∫ ∫ dx dz dy. (13.6.20)
−1 0 −√1−y 2

−1 ≤ y ≤ 0

The following theorem states two things that should make "common sense'' to us. First, using the triple integral to find volume of a
region D should always return a positive number; we are computing volume here, not signed volume. Secondly, to compute the
volume of a "complicated'' region, we could break it up into subregions and compute the volumes of each subregion separately,
summing them later to find the total volume.

THEOREM 126: Properties of Triple Integrals


Let D be a closed, bounded region in space, and let D and D be non-overlapping regions such that D = D
1 2 1 ⋃ D2 .
1. ∭ D
dV ≥ 0

2. ∭ D
dV = ∭
D1
dV + ∭
D2
dV .

We use this latter property in the next example.

Example 13.6.5: Finding the volume of a space region with triple integration

Find the volume of the space region D bounded by the coordinate planes, z = 1 − x/2 and z = 1 − y/4 , as shown in Figure
13.43(a). Set up the triple integrals that find the volume of D in all 6 orders of integration.

13.6.8 https://math.libretexts.org/@go/page/4239
Figure 13.43: The region D in Example 13.6.5 is shown in (a); in (b), it is collapsed onto the x-y plane.
Solution
Following the bounds--determining strategy of "surface to surface, curve to curve, and point to point,'' we can see that the most
difficult orders of integration are the two in which we integrate with respect to z first, for there are two "upper'' surfaces that
bound D in the z -direction. So we start by noting that we have
1 1
0 ≤z ≤1− x and 0 ≤z ≤1− y. (13.6.21)
2 4

We now collapse the region D onto the x-y axis, as shown in Figure 13.43(b). The boundary of D, the line from (0, 0, 1) to
(2, 4, 0), is shown in part (b) of the figure as a dashed line; it has equation y = 2x. (We can recognize this in two ways: one, in

collapsing the line from (0, 0, 1) to (2, 4, 0) onto the x-y plane, we simply ignore the z -values, meaning the line now goes
from (0, 0) to (2, 4). Secondly, the two surfaces meet where z = 1 − x/2 is equal to z = 1 − y/4 : thus
1 − x/2 = 1 − y/4 ⇒ y = 2x. )

We use the second property of Theorem 126 to state that

∭ dV = ∭ dV + ∭ dV , (13.6.22)
D D1 D2

where D and D are the space regions above the plane regions R and R , respectively. Thus we can say
1 2 1 2

1−x/2 1−y/4

∭ dV = ∬ (∫ dz) dA + ∬ (∫ dz) dA. (13.6.23)


D R1 0 R2 0

All that is left is to determine bounds of R and R , depending on whether we are integrating with order
1 2 dxdy or dy dx. We
give the final integrals here, leaving it to the reader to confirm these results.
dz dy dx :
$$

\begin{array}{ccccc}
& &\begin{array}{c}
0\leq z\leq 1-x/2\\
0\leq y\leq 2x\\
0\leq x\leq 2
\end{array}

&&
0 ≤ z ≤ 1 − y/4

2x ≤ y ≤ 4 (13.6.24)

0 ≤x ≤2

\\
\\
\iiint_D dV &=& \int_0^2\int_0^{2x}\int_0^{1-x/2} dz \, dy \, dx &+& \int_0^2\int_{2x}^4\int_0^{1-y/4} dz \, dy \, dx
\end{array}
\]

13.6.9 https://math.libretexts.org/@go/page/4239
dz dx dy :
$$

\begin{array}{ccccc}
& &\begin{array}{c}
0\leq z\leq 1-x/2\\
y/2\leq x\leq 2\\
0\leq y\leq 4
\end{array}

&&
0 ≤ z ≤ 1 − y/4

0 ≤ x ≤ y/2 (13.6.25)

0 ≤y ≤4

\\
\\
\iiint_D dV &=& \int_0^4\int_{y/2}^{2}\int_0^{1-x/2} dz \, dx \, dy &+& \int_0^4\int_{0}^{y/2}\int_0^{1-y/4} dz \, dx \, dy
\end{array}
\]
The remaining four orders of integration do not require a sum of triple integrals. In Figure 13.44 we show D collapsed onto the
other two coordinate planes. Using these graphs, we give the final orders of integration here, again leaving it to the reader to
confirm these results.

Figure 13.44: The region D in Example 13.6.5 is shown collapsed onto the x-z plane in (a); in (b), it is collapsed onto the y-z
plane.
dy dx dz :

0 ≤ y ≤ 4 − 4z
1 2−2z 4−4z
0 ≤ x ≤ 2 − 2z ⇒ ∫ ∫ ∫ dy dx dz (13.6.26)
0 0 0

0 ≤z ≤1

dydzdx :
$$

\begin{array}{cc}
\begin{array}{c}
0\leq y\leq 4-4z\\
0\leq z\leq 1-x/2\\
0\leq x\leq 2
\end{array}

&
\Rightarrow \int_0^2\int_{0}^{1-x/2}\int_0^{4-4z} dy \, dx \, dz

13.6.10 https://math.libretexts.org/@go/page/4239
\end{array}
\]
dx dy dz :
$$

\begin{array}{cc}
\begin{array}{c}
0\leq x\leq 2-2z\\
0\leq y\leq 4-4z\\
0\leq z\leq 1
\end{array}

&
\Rightarrow \int_0^1\int_{0}^{4-4z}\int_0^{2-2z} dx \, dy \, dz
\end{array}
\]
dx dz dy :
$$

\begin{array}{cc}
\begin{array}{c}
0\leq x\leq 2-2z\\
0\leq z\leq 1-y/4\\
0\leq y\leq 4
\end{array}

&
\Rightarrow \int_0^4\int_{0}^{1-y/4}\int_0^{2-2z} dx \, dz \, dy
\end{array}
\]

We give one more example of finding the volume of a space region.

Example 13.6.6: Finding the volume of a space region

Set up a triple integral that gives the volume of the space region D bounded by z = 2x + 2 and z = 6 − 2x
2 2
−y
2
. These
surfaces are plotted in Figure 13.45(a) and (b), respectively; the region D is shown in part (c) of the figure.

Figure 13.45: The region D is bounded by the surfaces shown in (a) and (b); D is shown in (c).
Solution
The main point of this example is this: integrating with respect to z first is rather straightforward; integrating with respect to x
first is not.
The order dz dy dx :
The bounds on z are clearly 2x + 2 ≤ z ≤ 6 − 2x − y . Collapsing D onto the x-y plane gives the ellipse shown in Figure
2 2 2

13.45(c). The equation of this ellipse is found by setting the two surfaces equal to each other:

13.6.11 https://math.libretexts.org/@go/page/4239
2
2 2 2 2 2 2
y
2x + 2 = 6 − 2x −y ⇒ 4x +y =4 ⇒ x + = 1. (13.6.27)
4

We can describe this ellipse with the bounds


− −−−− − − −−−− −
2 2
−√ 4 − 4x ≤ y ≤ √ 4 − 4x and − 1 ≤ x ≤ 1. (13.6.28)

Thus we find volume as


2 2 2
2x + 2 ≤ z ≤ 6 − 2x −y

− −−−− − − −−−− − 1 √4−4x2 6−2 x −y


2 2

2 2
−√ 4 − 4x ≤ y ≤ √ 4 − 4x ⇒ ∫
−1

2
∫ 2
2 x +2
dz dy dx . (13.6.29)
−√4−4x

−1 ≤ x ≤ 1

The order dydzdx :


Integrating with respect to y is not too difficult. Since the surface z = 2x + 2 is a cylinder whose directrix is the y -axis, it
2

does not create a border for y . The paraboloid z = 6 − 2x − y does; solving for y , we get the bounds
2 2

− −−−−−−− − − −−−−−−− −
2 2
−√ 6 − 2 x − z ≤ y ≤ √ 6 − 2 x − z . (13.6.30)

Collapsing D onto the x-z axes gives the region shown in Figure 13.46(a); the lower curve is from the cylinder, with equation
z = 2 x + 2 . The upper curve is from the paraboloid; with y = 0 , the curve is z = 6 − 2x . Thus bounds on z are
2 2

2
2x + 2 ≤ z ≤ 6 − 2x ; the bounds on x are −1 ≤ x ≤ 1 . Thus we have:
2

− −−−−−−− − − −−−−−−− −
−√ 6 − 2 x2 − z ≤ y ≤ √ 6 − 2 x2 − z
2
1 6−2x √6−2 x2 −z
2 2
2x + 2 ≤ z ≤ 6 − 2x ⇒ ∫ ∫ 2
∫ dydzdx. (13.6.31)
−1 2 x +2 2
−√6−2 x −z

−1 ≤ x ≤ 1

The order dx dz dy :
This order takes more effort as D must be split into two subregions. The two surfaces create two sets of upper/lower bounds in
terms of x; the cylinder creates bounds
−−−−−− −−−−−−
−√ z/2 − 1 ≤ x ≤ √ z/2 − 1 (13.6.32)

for region D and the paraboloid creates bounds


1

−−−−−−−−−−−−− −−−−−−−−−−−−−
2 2 2 2
−√ 3 − y /2 − z /2 ≤ x ≤ √ 3 − y /2 − z /2 (13.6.33)

for region D .
2

Figure 13.46: The region D in Example 13.6.6 is collapsed onto the x-z plane in (a); in (b), it is collapsed onto the y-z plane.
Collapsing D onto the y -z axes gives the regions shown in Figure 13.46(b). We find the equation of the curve z = 4 − y /2
2

by noting that the equation of the ellipse seen in Figure 13.45(c) has equation
−−−−−−−
2 2 2
x + y /4 = 1 ⇒ x = √ 1 − y /4 . (13.6.34)

Substitute this expression for x in either surface equation, z = 6 − 2x 2


−y
2
or z = 2x 2
+2 . In both cases, we find
1 2
z =4− y . (13.6.35)
2

13.6.12 https://math.libretexts.org/@go/page/4239
Region R , corresponding to D , has bounds
1 1

2
2 ≤ z ≤ 4 − y /2, −2 ≤ y ≤ 2 (13.6.36)

and region R , corresponding to D , has bounds


2 2

2 2
4 − y /2 ≤ z ≤ 6 − y , −2 ≤ y ≤ 2. (13.6.37)

Thus the volume of D is given by:


$$\int_{-2}^2\int_2^{4-y^2/2}\int_{-\sqrt{z/2-1}}^{\sqrt{z/2-1}} dx \, dz \, dy \ +\ \int_{-2}^2\int_{4-y^2/2}^{6-y^2}\int_{-
\sqrt{3-y^2/2-z^2/2}}^{\sqrt{3-y^2/2-z^2/2}} dx \, dz \, dy.\]

If all one wanted to do in Example 13.6.6 was find the volume of the region D, one would have likely stopped at the first
integration setup (with order dz dy dx ) and computed the volume from there. However, we included the other two methods 1) to
show that it could be done, "messy'' or not, and 2) because sometimes we "have'' to use a less desirable order of integration in order
to actually integrate.

Triple Integration and Functions of Three Variables


There are uses for triple integration beyond merely finding volume, just as there are uses for integration beyond "area under the
curve.'' These uses start with understanding how to integrate functions of three variables, which is effectively no different than
integrating functions of two variables. This leads us to a definition, followed by an example.

Definition 107 Iterated Integration, (Part II)

Let D be a closed, bounded region in space, over which g 1 (x) ,g


2 (x) ,f1 (x, ,
y) f2 (x, y) and h(x, y, z) are all continuous, and
let a and b be real numbers.
b g2 (x) f2 (x,y)
The iterated integral ∫ a

g1 (x)

f1 (x,y)
h(x, y, z)dz dy dx is evaluated as

$$\int_a^b\int_{g_1(x)}^{g_2(x)}\int_{f_1(x,y)}^{f_2(x,y)} h(x,y,z) dz \, dy \, dx =
\int_a^b\int_{g_1(x)}^{g_2(x)}\left(\int_{f_1(x,y)}^{f_2(x,y)} h(x,y,z) dz\right) dy \, dx.\]

Example 13.6.7: Evaluating a triple integral of a function of three variables


1 x 2x+3y
Evaluate ∫ 0

x2

x2 −y
(xy + 2xz)dz dy dx.

Solution
We evaluate this integral according to Definition 107.
1 x 2x+3y
∫ ∫ ∫ (xy + 2xz)dz dy dx
0 x2 x2 −y

1 x 2x+3y

=∫ ∫ (∫ (xy + 2xz)dz) dy dx
0 x2 x2 −y

1 x
2x+3y
2 ∣
=∫ ∫ ((xyz + x z ) ) dy dx
∣ 2
2 x −y
0 x

1 x
2 2 2 2
=∫ ∫ (xy(2x + 3y) + x(2x + 3y ) − (xy(x − y) + x(x − y ) ))dy dx
2
0 x

1 x
5 3 3 2 2
=∫ ∫ (−x + x y + 4x + 14 x y + 12x y )dy dx.
2
0 x

We continue as we have in the past, showing fewer steps.


1
7 7 6
7 5 4
=∫ (− x − 8x − x + 15 x )dx
0
2 2

281
= ≈ 0.836.
336

13.6.13 https://math.libretexts.org/@go/page/4239
We now know how to evaluate a triple integral of a function of three variables; we do not yet understand what it means. We build
up this understanding in a way very similar to how we have understood integration and double integration.
Let h(x, y, z) a continuous function of three variables, defined over some space region D. We can partition D into n rectangular--
solid subregions, each with dimensions Δx × Δy × Δz . Let (x , y , z ) be some point in the i subregion, and consider the
i i i i i i
th

product h(x , y , z )Δx Δy Δz . It is the product of a function value (that's the h(x , y , z ) part) and a small volume ΔV (that's
i i i i i i i i i i

the Δx Δy Δz part). One of the simplest understanding of this type of product is when h describes the density of an object, for
i i i

then h × volume = mass .


We can sum up all n products over D. Again letting ||ΔD|| represent the length of the longest diagonal of the n rectangular solids
in the partition, we can take the limit of the sums of products as ||ΔD|| → 0. That is, we can find
$$ S = \lim_{||\Delta D||\to 0} \sum_{i=1}^n h(x_i,y_i,z_i)\Delta V_i=\lim_{||\Delta D||\to 0} \sum_{i=1}^n h(x_i,y_i,z_i)\Delta
x_i\Delta y_i\Delta z_i.\]
While this limit has lots of interpretations depending on the function h , in the case where h describes density, S is the total mass of
the object described by the region D.
We now use the above limit to define the triple integral, give a theorem that relates triple integrals to iterated iteration, followed
by the application of triple integrals to find the centers of mass of solid objects.

Definition 108 Triple Integral


Let w = h(x, y, z) be a continuous function over a closed, bounded space region D, and let ΔD be any partition of D into n
rectangular solids with volume ΔV . The triple integral of h over D is
i

$$\iiint_Dh(x,y,z) dV = \lim_{||\Delta D||\to 0}\sum_{i=1}^n h(x_i,y_i,z_i)\Delta V_i.\]

Note: We previously showed how the summation of rectangles over a region R in the plane could be viewed as a double sum,
leading to the double integral. Likewise, we can view the sum ∑ h(x , y , z )Δx Δy Δz as a triple sum, n

i=1 i i i i i i

p n m

∑ ∑ ∑ h(xi , yj , zk )Δxi Δyj Δzk , (13.6.38)

k=1 j=1 i=1

which we evaluate as
p n m

∑ ( ∑ ( ∑ h(xi , yj , zk )Δxi ) Δyj ) Δzk . (13.6.39)

k=1 j=1 i=1

Here we fix a k value, which establishes the z -height of the rectangular solids on one "level'' of all the rectangular solids in the
space region D. The inner double summation adds up all the volumes of the rectangular solids on this level, while the outer
summation adds up the volumes of each level.
This triple summation understanding leads to the ∭ notation of the triple integral, as well as the method of evaluation shown in
D

Theorem 127.
The following theorem assures us that the above limit exists for continuous functions h and gives us a method of evaluating the
limit.

theorem 127 Triple Integration (Part II)


Let w = h(x, y, z) be a continuous function over a closed, bounded space region D, and let ΔD be any partition of D into n
rectangular solids with volume V . i

n
1. The limit lim ∑
||ΔD||→0 h(x , y , z )ΔV exists.
i=1 i i i i

2. If D is defined as the region bounded by the planes x = a and x = b , the cylinders y = g (x) and y = g (x) , and the 1 2

surfaces z = f (x, y) and z = f (x, y) , where a < b , g (x) ≤ g (x) and f (x, y) ≤ f (x, y) on D, then
1 2 1 2 1 2

b g2 (x) f2 (x,y)

∭ h(x, y, z)dV = ∫ ∫ ∫ h(x, y, z)dz dy dx. (13.6.40)


D a g (x) f (x,y)
1 1

We now apply triple integration to find the centers of mass of solid objects.

13.6.14 https://math.libretexts.org/@go/page/4239
Mass and Center of Mass
One may wish to review Section 13.4 for a reminder of the relevant terms and concepts.

Definition 109 Mass, Center of Mass of Solids


Let a solid be represented by a region D in space with variable density function δ(x, y, z).
1. The mass of the object is M = ∭ dm = ∭
D D
δ(x, y, z)dV .
2. The moment about the x-y plane is M = ∭ xy D
zδ(x, y, z)dV .
3. The moment about the x-z plane is M = ∭ xz
D
yδ(x, y, z)dV .

4. The moment about the y -z plane is M = ∭yz


D
xδ(x, y, z)dV .

5. The center of mass of the object is


Myz Mxz Mxy
¯¯
¯ ¯
¯¯ ¯
¯¯
(x, y , z ) = ( , , ). (13.6.41)
M M M

Example 13.6.8: Finding the center of mass of a solid

Find the mass and center of mass of the solid represented by the space region bounded by the coordinate planes and
z = 2 − y/3 − 2x/3 , shown in Figure 13.47, with constant density δ(x, y, z) = 3 gm/cm . (Note: this space region was used
3

in Example 13.6.2.)

Figure 13.47: Finding the center of mass of this solid in Example 13.6.8.
Solution
We apply Definition 109. In Example 13.6.2, we found bounds for the order of integration dz dy dx to be
0 ≤ z ≤ 2 − y/3 − 2x/3 , 0 ≤ y ≤ 6 − 2x and 0 ≤ x ≤ 3 . We find the mass of the object:

M =∭ δ(x, y, z)dV
D

3 6−2x 2−y/3−2x/3

=∫ ∫ ∫ (3)dz dy dx
0 0 0

3 6−2x 2−y/3−2x/3

=3∫ ∫ ∫ dz dy dx
0 0 0

= 3(6) = 18gm.

The evaluation of the triple integral is done in Example 13.6.2, so we skipped those steps above. Note how the mass of an
object with constant density is simply "density×volume.''
We now find the moments about the planes.

13.6.15 https://math.libretexts.org/@go/page/4239
Mxy = ∭ 3zdV
D

3 6−2x 2−y/3−2x/3

=∫ ∫ ∫ (3z)dz dy dx
0 0 0

3 6−2x
3 2
=∫ ∫ (2 − y/3 − 2x/3 ) dy dx
0 0
2

3
4 3
=∫ − (x − 3 ) dx
0
9

= 9.

We omit the steps of integrating to find the other moments.

Myz = ∭ 3xdV
D

27
= .
2

Mxz = ∭ 3ydV
D

= 27.

The center of mass is $$\big(\overline{x},\overline{y},\overline{z}\big) = \left(\frac{27/2}{18},\frac{27}{18},\frac{9}


{18}\right) = \big(0.75,1.5,0.5\big).\]

Example 13.6.9: Finding the center of mass of a solid

Find the center of mass of the solid represented by the region bounded by the planes z = 0 and z = −y and the cylinder
x + y = 1 , shown in Figure 13.48, with density function δ(x, y, z) = 10 + x + 5y − 5z . (Note: this space region was used
2 2 2

in Example 13.6.3.)

Figure 13.48: Finding the center of mass of this solid in Example 13.6.9.
Solution
As we start, consider the density function. It is symmetric about the y -z plane, and the farther one moves from this plane, the
denser the object is. The symmetry indicates that x should be 0.
¯¯
¯

As one moves away from the origin in the y or z directions, the object becomes less dense, though there is more volume in
these regions.
Though none of the integrals needed to compute the center of mass are particularly hard, they do require a number of steps. We
emphasize here the importance of knowing how to set up the proper integrals; in complex situations we can appeal to
technology for a good approximation, if not the exact answer. We use the order of integration dz dy dx , using the bounds found
in Example 13.6.4. (As these are the same for all four triple integrals, we explicitly show the bounds only for M .)

13.6.16 https://math.libretexts.org/@go/page/4239
2
M =∭ (10 + x + 5y − 5z)dV
D

1 0 −y
2
=∫ ∫ ∫ (10 + x + 5y − 5z)dV
2
−1 −√1−x 0

64 15π
= − ≈ 3.855.
5 16

2
Myz =∭ x(10 + x + 5y − 5z)dV
D

= 0.

2
Mxz =∭ y(10 + x + 5y − 5z)dV
D

61π
=2− ≈ −1.99.
48

2
Mxy =∭ z(10 + x + 5y − 5z)dV
D

61π 10
= − ≈ 0.885.
96 9

Note how M = 0 , as expected. The center of mass is


yz

$$\big(\overline{x},\overline{y},\overline{z}\big) = \left(0,\frac{-1.99}{3.855},\frac{0.885}{3.855}\right) \approx


\big(0,-0.516, 0.230\big).\]

As stated before, there are many uses for triple integration beyond finding volume. When h(x, y, z) describes a rate of change
function over some space region D, then ∭ h(x, y, z)dV gives the total change over D. Our one specific example of this was
D

computing mass; a density function is simply a "rate of mass change per volume'' function. Integrating density gives total mass.
While knowing how to integrate is important, it is arguably much more important to know how to set up integrals. It takes skill to
create a formula that describes a desired quantity; modern technology is very useful in evaluating these formulas quickly and
accurately.
This chapter investigated the natural follow--on to partial derivatives: iterated integration. We learned how to use the bounds of a
double integral to describe a region in the plane using both rectangular and polar coordinates, then later expanded to use the bounds
of a triple integral to describe a region in space. We used double integrals to find volumes under surfaces, surface area, and the
center of mass of lamina; we used triple integrals as an alternate method of finding volumes of space regions and also to find the
center of mass of a region in space.
Integration does not stop here. We could continue to iterate our integrals, next investigating "quadruple integrals'' whose bounds
describe a region in 4--dimensional space (which are very hard to visualize). We can also look back to "regular'' integration where
we found the area under a curve in the plane. A natural analogue to this is finding the "area under a curve,'' where the curve is in
space, not in a plane. These are just two of many avenues to explore under the heading of "integration.''

This page titled 13.6: Volume Between Surfaces and Triple Integration is shared under a CC BY-NC 3.0 license and was authored, remixed,
and/or curated by Gregory Hartman et al. via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit
history is available upon request.

13.6.17 https://math.libretexts.org/@go/page/4239
13.E: Applications of Multiple Integration (Exercises)
13.1: Iterated Integrals and Area
Terms and Concepts
1. When integrating f x (x, y) with respect to x, the constant of integration C is really which: C (x) or C (y)? What does this mean?
2. Integrating an integral is called _________ __________.
3. When evaluating an iterated integral, we integrate from _______ to ________, then from _________ to __________.
b g2 (x)

4. One understanding of an iterated integral is that ∫ ∫ dy dx gives the _______ of a plane region.
a g1 (x)

Problems
In Exercises 5-10, evaluate the integral and subsequent iterated integral.
5.
5

(a) ∫ (6 x
2
+ 4xy − 3 y ) dy
2

2
2 5

(b) ∫ ∫ (6 x
2
+ 4xy − 3 y ) dy dx
2

−3 2

6.
π

(a) ∫ (2x cos y + sin x) dx


0
π/2 π

(b) ∫ ∫ (2x cos y + sin x) dx dy


0 0

7.
x

(a) ∫ 2
(x y − y + 2) dy
1
2 x

(b) ∫ ∫
2
(x y − y + 2) dy dx
0 1

8.
2
y

(a) ∫ (x − y) dx
y
2
1 y

(b) ∫ ∫ (x − y) dx dy
−1 y

9.
y

(a) ∫ (cos x sin y) dx


0
π y

(b) ∫ ∫ (cos x sin y) dx dy


0 0

10.
x
1
(a) ∫ (
2
) dy
0 1 +x
2 x
1
(b) ∫ ∫ (
2
) dy dx
1 0 1 +x

In Exercises 11-16, a graph of a planar region R is given. Give the iterated integrals, with both orders of integration dy dx
and dx dy, that give the area of R. Evaluate one of the iterated integrals to find the area.

13.E.1 https://math.libretexts.org/@go/page/10008
11.

12.

13.

14.

15.

16.

In Exercises 17-22, iterated integrals are given that compute the area of a region R in the xy-plane. Sketch the region R,
and give the iterated integral(s) that give the area of R with the opposite order of integration.

13.E.2 https://math.libretexts.org/@go/page/10008
2
2 4−x

17. ∫ ∫ dy dx
−2 0

2
1 5−5x

18. ∫ ∫ dy dx
0 5−5x

2
2 2 √4−y

19. ∫ ∫ dx dy
−2 0

3 √9−x2

20. ∫ ∫ dy dx
2
−3 −√9−x

1 √y 4 √y

21. ∫ ∫ dx dy + ∫ ∫ dx dy
0 −√y 1 y−2

1 (1−x)/2

22. ∫ ∫ dy dx
−1 (x−1)/2

13.2: Double Integration and Volume


Terms and Concepts
1. An integral can be interpreted as giving the signed area over an interval; a double integral can be interpreted as giving the signed
________ over a region.
b g2 (x) b g2 (y)
2. Explain why the following statement is false: "Fubini's Theorem states that ∫ a

g1 (x)
f (x, y) dy dx = ∫
a

g1 (y)
."
f (x, y) dx dy

3. Explain why if f (x, y) > 0 over a region R, then ∫ ∫ R


f (x, y) dA > 0 .
4. If ∫ ∫ R
f (x, y)dA = ∫ ∫
R
g(x, y) dA , does this imply f (x, y) = g(x, y)?

Problems
In Exercises 5-10,
(a) Evaluate the given iterated integral, and
(b) rewrite the integral using the other order of integration.
2 1
5. ∫1

−1
(
x

y
+ 3) dx dy

π/2 π
6. ∫−π/2

0
(sin x cos y \,dy\,dx\)
4 −x/2+2
7. ∫0

0
(3 x
2
− y + 2) dy dx

3 3
8. ∫1

y
2
(x y − x y ) dx dy
2

2 √1−y
9. ∫0
1∫
−√1−y
(x + y + 2) dx dy

9 √3
10. ∫ 0

y/3
(x y ) dx dy
2

In Exercises 11-18:
(a) Sketch the region R given by the problem.
(b) Set up the iterated integrals, in both orders, that evaluate the given double integral for the described region R.
(c) Evaluate one of the iterated integrals to find the signed volume under the surface z = f (x, y) over the region R.
11. ∫ ∫ R
2
x y dA , where R is bounded by y = √−
x and y = x .
2

12. ∫ ∫ R
2
x y dA , where R is bounded by y = √−
x and y = x .
3 3

13. ∫ ∫ R
x
2
−y
2
dA , where R is the rectangle with corners (−1, −1), (1, −1), (1, 1) and (−1, 1).
14. ∫ ∫ R
ye
x
dA , where R is bounded by x = 0, x =y
2
and y = 1 .
15. ∫ ∫ R
(6 − 3x − 2y) dA , where R is bounded by x = 0, y = 0 and 3x + 2y = 6 .

13.E.3 https://math.libretexts.org/@go/page/10008
16. ∫ ∫ R
e
y
dA , where R is bounded by y = ln x and y = 1

e−1
(x − 1) .
17. ∫ ∫ R
3
(x y − x) dA , where R is the half of the circle x 2
+y
2
=9 in the first and second quadrants.
18. ∫ ∫ R
(4 − sy) dA , where R is bounded by y = 0, y = x/e and y = ln x .
In Exercises 19-22, state why it is difficult/impossible to integrate the iterated integral in the given order of integration.
Change the order of integration and evaluate the new iterated integral.
2
4 2
19. ∫ 0

y/2
e
x
dx dy

√π/2 √π/2
20. ∫ 0

x
cos(y ) dy dx
2

1 1 2y
21. ∫ 0

y x +y
2 2
dx dy

2
1 2 x tan y
22. ∫ −1

1 1+ln y
dy dx

In Exercises 23-26, find the average value of f over the region R. Notice how these functions and regions are related to the
iterated integrals given in Exercises 5-8.
23. f (x, y) = x

y
+3 ; R is the rectangle with opposite corners (−1, 1) and (1, 2).
24. f (x, y) = sin x cos y; R is bounded by x = 0, x = π, y = −π/2 and y = π/2 .
25. f (x, y) = 3x 2
−y +2 ; R is bounded by the lines y = 0, y = 2 − x/2 and x = 0 .
26. f (x, y) = x 2
y − xy
2
; R is bounded by y = x, y = 1 and x = 3 .

13.3: Double Integration with Polar Coordinates


Terms and Concepts
1. When evaluating ∫ ∫ R
f (x, y) dA using polar coordinates, f (x, y) is replaced with _______ and dA is replaced with _______.
2. Why would one be interested in evaluating a double integral with polar coordinates?

Problems
In Exercises 3-10, a function f (x, y) is given and a region R of the x-y plane is described. Set up and evaluate
∫ ∫ f (x, y) dA .
R

3. f (x, y) = 3x − y + 4 ; R is the region enclosed by the circle x 2


+y
2
=1 .
4. f (x, y) = 4x + 4y ; R is the region enclosed by the circle x 2
+y
2
=4 .
5. f (x, y) = 8 − y ; R is the region enclosed by the circles with polar equations r = cos θ and r = 3 cos θ .
6. f (x, y) = 4; R is the region enclosed by the petal of the rose curve r = sin(2θ) in the first quadrant.
7. f (x, y) = ln(x 2
+y )
2
; R is the annulus enclosed by the circles \(x^2+y^2=1\text{ and }x^2+y^2=4.
8. f (x, y) = 1 − x 2
−y
2
; R is the region enclosed by the circle x 2
+y
2
=1 .
9. f (x, y) = x 2
−y
2
; R is the region enclosed by the circle x 2
+y
2
= 36 in the first and fourth quadrants.
10. f (x, y) = (x − y)/(x + y) ; R is the region enclosed by the lines y = x, y = 0 and the circle x 2
+y
2
=1 in the first quadrant.
In Exercises 11-14, an iterated integral in rectangular coordinates is given. Rewrite the integral using polar coordinates and
evaluate the new double integral.
5 √25−x2 −−−−− −
11. ∫ 0

−√25−x
2
√x2 + y 2 dy dx

4 0
12. ∫ −4

−√16−y
2
(2y − x)dx dy

2
2 √8−y
13. ∫ 0

y
(x + y) dx dy

−1 √4−x2 1 √4−x2 2 √4−x2


14. ∫ −2

0
(x + 5)dy dx + ∫
−1

√1−x2
(x + 5) dy dx + ∫
1

0
(x + 5) dy dx

13.E.4 https://math.libretexts.org/@go/page/10008
In Exercises 15-16, special double integrals are presented that are especially well suited for evaluation in polar coordinates.
15. Consider ∫ ∫ e
2 2
−( x +y )
dA.
R

(a) Why is this integral difficult to evaluate in rectangular coordinates, regardless of the region R?
(b) Let R be the region bounded by the circle of radius a centered at the origin. Evaluate the double integral using polar
coordinates.
2 2

(c) Take the limit of your answer from (b), as a → ∞ . What does this imply about the volume under the surface of e over −( x +y )

the entire x-y plane?


16. The surface of a right circular cone with height h and base radius a can be described by the equation
−−−−−−
2
y
, where the tip of the cone lies at and the circular base lies in the x-y plane, centered at the
2
x
f (x, y) = h − h √ + (0, 0, h)
a2 a2

origin.
Confirm that the volume of a right circular cone with height h and base radius a is V =
1

3
2
πa h by evaluating ∫ ∫
R
f (x, y) dA in
polar coordinates.

13.4: Center of Mass


Terms and Concepts
1. Why is it easy to use "mass" and "weight" interchangeably, even though they are different measures?
2. Given a point (x, y), the value of x is a measure of distance from the _________-axis.
3. We can think of ∫ ∫ R
dm as meaning "sum up lots of ________."
4. What is a "discrete planar system?"
5. Why does M use ∫ ∫ x R
yδ(x, y) dA instead of ∫ ∫ R
xδ(x, y) dA ; that is, why do we use "y" and not "x"?
6. Describe a situation where the center of mass of a lamina does not lie within the region of the lamina itself.

Problems
In Exercises 7-10, point masses are given along a line or in the plane. Find the center of mass ¯¯
x
¯
or ¯¯
¯ ¯
¯
(x, y )
¯
, as appropriate.
(All masses are in grams and distances are in cm.)
7. m 1 = 4 at x = 1; m2 = 3 at x = 3; m3 = 5 at x = 10

8. m 1 = 2 at x = −3; m2 = 2 at x = −1; m3 = 3 at x = 0; m4 = 3 at x = 7

9. m 1 = 2 at (−2, 2); m2 = 2 at (2, −2); m3 = 20 at (0, 4)

10. m 1 = 1 at (−1, 1); m2 = 2 at (−1, 1); m3 = 2 at (1, 1); m4 = 1 at (1, −1)

In Exercises 11-18, find the mass/weight of the lamina described by the region R in the plane and its density function δ(x, y).
11. R is the rectangle with corners (1, −3), (1, 2), (7, 2) and (7, −3); δ(x, y) = 5gm/cm 2

12. R is the rectangle with corners (1, −3), (1, 2), (7, 2) and (7, −3); δ(x, y) = (x + y )gm/cm 2 2

13. R is the triangle with corners (−1, 0), (1, 0), and (0, 1); δ(x, y) = 2 lb/in 2

14. R is the triangle with corners (0, 0), (1, 0), and (0, 1); δ(x, y) = (x
2
+y
2
+ 1) lb/in 2

15. R is the circle centered at the origin with radius 2; δ(x, y) = (x + y + 4) kg/m 2

−−−−− −
16. R is the circle sector bounded by x 2
+y
2
= 25 in the first quadrant; δ(x, y) = (√x 2
+ y 2 + 1) kg/m 2

17. R is the annulus in the first and second quadrants bounded by x 2


+y
2
= 9 and x
2
+y
2
= 36; δ(x, y) = 4 lb/ft 2

−−−−−−
18. R is the annulus in the first and second quadrants bounded by x 2
+y
2
= 9 and x
+
y
2 2
= 36; δ(x, y) = √x + y
2
lb/ft 2

In Exercises 19-26, find the center of mass of the lamina described by the region R in the plane and its density function
δ(x, y).

Note: these are the same lamina as in Exercises 11-18.


19. R is the rectangle with corners (1, −3), (1, 2), (7, 2) and (7, −3); δ(x, y) = 5gm/cm 2

13.E.5 https://math.libretexts.org/@go/page/10008
20. R is the rectangle with corners (1, −3), (1, 2), (7, 2) and (7, −3); δ(x, y) = (x + y )gm/cm 2 2

21. R is the triangle with corners (−1, 0), (1, 0), and (0, 1); δ(x, y) = 2 lb/in 2

22. R is the triangle with corners (0, 0), (1, 0), and (0, 1); δ(x, y) = (x
2
+y
2
+ 1) lb/in 2

23. R is the circle centered at the origin with radius 2; δ(x, y) = (x + y + 4) kg/m 2

−−−−−−
24. R is the circle sector bounded by x 2
+y
2
= 25 in the first quadrant; δ(x, y) = (√x 2
+y
2
+ 1) kg/m 2

25. R is the annulus in the first and second quadrants bounded by x 2


+y
2
= 9 and x
2
+y
2
= 36; δ(x, y) = 4 lb/ft2

−−−−−−
26. R is the annulus in the first and second quadrants bounded by x 2
+y
2
= 9 and x
+
y
2 2
= 36; δ(x, y) = √x + y
2
lb/ft
2

The moment of inertia i is a measure of the tendency of lamina to resist rotating about an axis or continue to rotate about
an axis. i is the moment of inertia about the x-axis, i is the moment of inertia about the x-axis, and i is the moment of
x x o

inertia about the origin.These are computed as follows:


2
ix = ∫ ∫ y dm
R
2
iy = ∫ ∫ x dm
R
2 2
io = ∫ ∫ (x + y ) dm
R

In Exercises 27-30, a lamina corresponding to a planar region R is given with a mass of 16 units. For each, compute ix , iy

and i .
o

27. R is the 4 x 4 square with corners (−2, −2) and (2, 2) with density δ(x, y) = 1.
28. R is the 8 x 2 rectangle with corners (−4, −1) and (4, 1) with density δ(x, y) = 1.
29. R is the 4 x 2 rectangle with corners (−2, −1) and (2, 1) with density δ(x, y) = 2.
30. R is the circle with radius 2 centered at the origin with density δ(x, y) = 4/π.

13.5: Surface Area


Terms and Concepts
1. "Surface area" is analogous to what previously studied concept?
2. To approximate the area of a small portion of a surface, we computed the area of its ______ plane.
3. We interpret ∫ ∫ R
dS as "sum up lots of little _______ ________."
4. Why is it important to know how to set up a double integral to compute surface area, even if the resulting integral is hard to
evaluate?
5. Why do z = f (x, y) and z = g(x, y) = f (x, y) + h , for some real number h, have the same surface area over a region R?
6. Let z = f (x, y) and z = g(x, y) = 2f (x, y) . Why is the surface area of g over a region R not twice the surface area of f over
R?

Problems
In Exercises 7-10, set up the iterated integral that computes the surfaces area of the given surface over the region R.
7. f (x, y) = sin x cos y; R is the rectangle with bounds 0 ≤ x ≤ 2π , 0 ≤ y ≤ 2π .

13.E.6 https://math.libretexts.org/@go/page/10008
8. f (x, y) = x2 +y 2 +1
1
; R is the circle x2
+y
2
=9 .

9. f (x, y) = x 2
−y ;
2
R is the rectangle with opposite corners (−1, −1) and 1, 1).

10. f (x, y) = x
2
1
; R is the rectangle bounded by −5 ≤ x ≤ 5 and 0 ≤ y ≤ 1 .
e +1

In Exercises 11-19, find the area of the given surface over the region R.
11. f (x, y) = 3x − 7y + 2; R is the rectangle with opposite corners (−1, 0) and (1, 3).
12. f (x, y) = 2x + 2y + 2; R is the triangle with corners (0, 0), (1, 0) and (0, 1).
13. f (x, y) = x 2
+y
2
+ 10; R is the circle x 2
+y
2
= 16 .
14. f (x, y) = −2x + 4y 2
+ 7 over R , the triangle bounded by y = −x, y = x, 0 ≤ y ≤ 1 .
15. f (x, y) = x 2
+y over R, the triangle bounded by y = 2x, y = 0 and x = 2 .
16. f (x, y) = 2

3
x
3/2
over R, the rectangle with opposite corners (0, 0) and (1, 1).
−−−−−−
17. f (x, y) = 10 − 2√x + y over R, the circle
2 2 2
x +y
2
= 25 . (This is the cone with height 10 and base radius 5; be sure to
compare your result with the known formula.)
−−−−−−−−− −
18. Find the surface area of the sphere with radius 5 by doubling the surface area of f (x, y) = √25 − x 2
−y
2
over R, the circle
x + y = 25 . (Be sure to compare your result with the known formula.)
2 2

19. Find the surface area of the ellipse formed by restricting the plane f (x, y) = cx + dy + h to the region R, the circle
x + y = 1 , where c, d and h are some constants. Your answer should be given in terms of c and d; why does the value of h not
2 2

matter?

13.6: Volume Between Surfaces and Triple Integration


Terms and Concepts
1. The strategy for establishing bounds for triple integrals is "________ to ________, _________ and __________ to _______."
2. Give an informal interpretation of what " ∫ ∫ ∫ D
dV " means.
3. Give two uses of triple integration.

13.E.7 https://math.libretexts.org/@go/page/10008
4. If an object has a constant density δ and a volume V, what is its mass?

Problems
In Exercises 5-8, two surfaces f (x, y) and f (x, y) and a region R in the xy-plane are given. Set up and evaluate the triple
1 2

integral that represents the volume between these surfaces over R.


5. f (x, y) = 8 − x − y , f (x, y) = 2x + y;
1
2 2
2

R is the square with corners (−1, −1) and (1, 1).

6. f (x, y) = x + y , f (x, y) = −x − y ;
1
2 2
2
2 2

R is the square with corners (0, 0) and (2, 3).

7. f (x, y) = sin x cos y, f (x, y) = cos x sin y + 2;


1 2

R is the triangle with corners (0, 0), (π, 0) and (π, π).

8. f (x, y) = 2x
1
2
+ 2y
2
+ 3, f2 (x, y) = 6 − x
2 2
−y ;

R is the circle x .
2 2
+y =1

In Exercises 9-16, a domain D is described by its bounding surfaces, along with a graph. Set up the triple integrals that give
the volume of D in all 6 orders of integration, and find the volume of D by evaluating the indicated triple integral.
9. D is bounded by the coordinate planes and z = 2 − 2x/3 − 2y .
Evaluate the triple integral with order dz dy dx.

10. D is bounded by the planes y = 0, y = 2, x = 1, z = 0 and z = (2 − x)/2 .


Evaluate the triple integral with order dx dy dz.

11. D is bounded by the planes x = 0, x = 2, z = −y and by z = y 2


/2 .
Evaluate the triple integral with order dy dz dx.

−− −−−− −
12. D is bounded by the planes z = 0, y = 9, x = 0 and by z = √y 2
− 9 x2 .
Do not evaluate any triple integral.

13.E.8 https://math.libretexts.org/@go/page/10008
13. D is bounded by the planes x = 2, y = 1, z = 0 and z = 2x + 4y − 4 .
Evaluate the triple integral with order dx dy dz.

14. D is bounded by the plane z = 2y and by y = 4 − x . 2

Evaluate the triple integral with order dz dy dx.

15. D is bounded by the coordinate planes and y = 1 − x and y = 1 − z .


2 2

Do not evaluate any triple integral. Which order is easier to evaluate: dz dy dx or dy dz dx? Explain why.

16. D is bounded by the coordinate planes and by z = 1 − y/3 and z = 1 − x .


Evaluate the triple integral with order dx dy dz.

In Exercises 17-20, evaluate the triple integral.


π/2 π π
17. ∫−π/2

0

0
(cos x sin y sin z)dz dy dx

13.E.9 https://math.libretexts.org/@go/page/10008
1 x x+y
18. ∫0

0

0
(x + y + z)dz dy dx

π 1 z
19. ∫0

0

0
(sin(yz))dx dy dz

2 3 2
π x y
20. ∫π

x

−y
2 (cos x sin y sin z)dz dy dx

In Exercises 21-24, find the center of mass of the solid represented by the indicated space region D with density function
δ(x, y, z).

21. D is bounded by the coordinate planes and z = 2 − 2x/3 − 2y ; δ(x, y, z) = 10 g/cm . 3

(Note: this is the same region as used in Exercise 9.)


22. D is bounded by the planes y = 0, y = 2, x = 1, z = 0 and z = (3 − x)/2 ; δ(x, y, z) = 2 g/cm . 3

(Note: this is the same region as used in Exercise 10.)


23. D is bounded by the planes x = 2, y = 1, z = 0 and z = 2x + 4y − 4 ; δ(x, y, z) = x lb/in . 2 3

(Note: this is the same region as used in Exercise 13.)


24. D is bounded by the planes z = 2y and by y = 4 − x . δ(x, y, z) = y lb/in .
2 2 3

(Note: this is the same region as used in Exercise 14.)

13.E: Applications of Multiple Integration (Exercises) is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by
LibreTexts.

13.E.10 https://math.libretexts.org/@go/page/10008
CHAPTER OVERVIEW
14: Appendix
Topic hierarchy
14.1: Section 1-
14.2: Section 2-
14.3: Section 3-
14.4: Section 4-
14.5: Section 5-
14.6: Section 6-

Contributors and Attributions


Gregory Hartman (Virginia Military Institute). Contributions were made by Troy Siemers and Dimplekumar Chalishajar of
VMI and Brian Heinold of Mount Saint Mary's University. This content is copyrighted by a Creative Commons Attribution -
Noncommercial (BY-NC) License. http://www.apexcalculus.com/

This page titled 14: Appendix is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et al. via
source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

1
14.1: Section 1-

Your page has been created!


Remove this content and add your own.

Edit page
Click the Edit page button in your user bar. You will see a suggested structure for your content. Add your content and hit
Save.

Tips:

Drag and drop


Drag one or more image files from your computer and drop them onto your browser window to add them to your page.

Classifications
Tags are used to link pages to one another along common themes. Tags are also used as markers for the dynamic organization
of content in the CXone Expert framework.

Working with templates


CXone Expert templates help guide and organize your documentation, making it flow easier and more uniformly. Edit
existing templates or create your own.

Visit for all help topics.

This page titled 14.1: Section 1- is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et al.
via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

14.1.1 https://math.libretexts.org/@go/page/4241
14.2: Section 2-

Your page has been created!


Remove this content and add your own.

Edit page
Click the Edit page button in your user bar. You will see a suggested structure for your content. Add your content and hit
Save.

Tips:

Drag and drop


Drag one or more image files from your computer and drop them onto your browser window to add them to your page.

Classifications
Tags are used to link pages to one another along common themes. Tags are also used as markers for the dynamic organization
of content in the CXone Expert framework.

Working with templates


CXone Expert templates help guide and organize your documentation, making it flow easier and more uniformly. Edit
existing templates or create your own.

Visit for all help topics.

This page titled 14.2: Section 2- is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et al.
via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

14.2.1 https://math.libretexts.org/@go/page/4242
14.3: Section 3-

Your page has been created!


Remove this content and add your own.

Edit page
Click the Edit page button in your user bar. You will see a suggested structure for your content. Add your content and hit
Save.

Tips:

Drag and drop


Drag one or more image files from your computer and drop them onto your browser window to add them to your page.

Classifications
Tags are used to link pages to one another along common themes. Tags are also used as markers for the dynamic organization
of content in the CXone Expert framework.

Working with templates


CXone Expert templates help guide and organize your documentation, making it flow easier and more uniformly. Edit
existing templates or create your own.

Visit for all help topics.

This page titled 14.3: Section 3- is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et al.
via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

14.3.1 https://math.libretexts.org/@go/page/4243
14.4: Section 4-

Your page has been created!


Remove this content and add your own.

Edit page
Click the Edit page button in your user bar. You will see a suggested structure for your content. Add your content and hit
Save.

Tips:

Drag and drop


Drag one or more image files from your computer and drop them onto your browser window to add them to your page.

Classifications
Tags are used to link pages to one another along common themes. Tags are also used as markers for the dynamic organization
of content in the CXone Expert framework.

Working with templates


CXone Expert templates help guide and organize your documentation, making it flow easier and more uniformly. Edit
existing templates or create your own.

Visit for all help topics.

This page titled 14.4: Section 4- is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et al.
via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

14.4.1 https://math.libretexts.org/@go/page/4244
14.5: Section 5-

Your page has been created!


Remove this content and add your own.

Edit page
Click the Edit page button in your user bar. You will see a suggested structure for your content. Add your content and hit
Save.

Tips:

Drag and drop


Drag one or more image files from your computer and drop them onto your browser window to add them to your page.

Classifications
Tags are used to link pages to one another along common themes. Tags are also used as markers for the dynamic organization
of content in the CXone Expert framework.

Working with templates


CXone Expert templates help guide and organize your documentation, making it flow easier and more uniformly. Edit
existing templates or create your own.

Visit for all help topics.

This page titled 14.5: Section 5- is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et al.
via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

14.5.1 https://math.libretexts.org/@go/page/4245
14.6: Section 6-

Your page has been created!


Remove this content and add your own.

Edit page
Click the Edit page button in your user bar. You will see a suggested structure for your content. Add your content and hit
Save.

Tips:

Drag and drop


Drag one or more image files from your computer and drop them onto your browser window to add them to your page.

Classifications
Tags are used to link pages to one another along common themes. Tags are also used as markers for the dynamic organization
of content in the CXone Expert framework.

Working with templates


CXone Expert templates help guide and organize your documentation, making it flow easier and more uniformly. Edit
existing templates or create your own.

Visit for all help topics.

This page titled 14.6: Section 6- is shared under a CC BY-NC 3.0 license and was authored, remixed, and/or curated by Gregory Hartman et al.
via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon request.

14.6.1 https://math.libretexts.org/@go/page/4246
Index
A curvature F
absolute convergence 11.5: The Arc Length Parameter and Curvature factorial
8.5: Alternating Series and Absolute Convergence curve sketching 8.1: Sequences
acceleration 3.5: Curve Sketching FINDING POSITION USING
11.3: The Calculus of Motion cusps
VELOCITY
Alternating Harmonic Series 9.2: Parametric Equations
5.2: The Definite Integral
8.5: Alternating Series and Absolute Convergence cylinders
First Derivative Test
alternating series 10.1: Introduction to Cartesian Coordinates in Space
3.3: Increasing and Decreasing Functions
8.5: Alternating Series and Absolute Convergence fluids
antiderivative D 7.6: Fluid Forces
5.1: Antiderivatives and Indefinite Integration decreasing on the interval I fundamental theorem of calculus
arc length 3.3: Increasing and Decreasing Functions
5.4: The Fundamental Theorem of Calculus
7.4: Arc Length and Surface Area definite integral
9.5: Calculus and Polar Functions
11.5: The Arc Length Parameter and Curvature
5.2: The Definite Integral
G
Derivative of Inverse Functions
Arc length of a parametric curve generalized chain rule
2.7: Derivatives of Inverse Functions
9.3: Calculus and Parametric Equations 2.5: The Chain Rule
Derivative of Inverse Trigonometric
arc length parameter generalized power rule
Functions 2.5: The Chain Rule
11.5: The Arc Length Parameter and Curvature
2.7: Derivatives of Inverse Functions
Derivative of logarithmic functions H
B 2.7: Derivatives of Inverse Functions
bisection method Higher Order Derivatives
Difference Rule 2.3: Basic Differentiation Rules
1.5: Continuity
5.1: Antiderivatives and Indefinite Integration
bounds of integration. Hyperbolas
Differentiability (two variables) 9.1: Conic Sections
5.2: The Definite Integral
12.4: Differentiability and the Total Differential
hyperbolic cosine
differentials
C 4.4: Differentials
6.6: Hyperbolic Functions
Cartesian Coordinate Plane hyperbolic functions
Differentiation Rules 6.6: Hyperbolic Functions
10.1: Introduction to Cartesian Coordinates in Space 2.3: Basic Differentiation Rules
Cartesian Coordinates in Space hyperbolic sine
Directional Derivatives 6.6: Hyperbolic Functions
10.1: Introduction to Cartesian Coordinates in Space 12.6: Directional Derivatives
center of mass directrix
13.4: Center of Mass
I
9.1: Conic Sections
chain rule disk method implicit differentiation
2.5: The Chain Rule 2.6: Implicit Differentiation
7.2: Volume by Cross-Sectional Area- Disk and
4.2: Related Rates
comparison test Washer Methods
12.5: The Multivariable Chain Rule
8.3: Integral and Comparison Tests distance IMPROPER INTEGRAL
Concavity 10.1: Introduction to Cartesian Coordinates in Space
6.8: Improper Integration
3.4: Concavity and the Second Derivative Distance between a Plane and a Point increasing on the interval I
9.3: Calculus and Parametric Equations 10.6: Planes
3.3: Increasing and Decreasing Functions
Concavity (parametric equations) double integral
9.3: Calculus and Parametric Equations
independent variable
13.2: Double Integration and Volume
5.1: Antiderivatives and Indefinite Integration
CONIC SECTIONS DOUBLE INTEGRAL IN POLAR
10.1: Introduction to Cartesian Coordinates in Space
indeterminate forms
COORDINATES 6.7: L'Hopital's Rule
conics
9.1: Conic Sections
13.3: Double Integration with Polar Coordinates infinite series
8.2: Infinite Series
constant multiple rule E
5.1: Antiderivatives and Indefinite Integration
inflection point
Constant Rule eccentricity 3.4: Concavity and the Second Derivative

2.3: Basic Differentiation Rules


9.1: Conic Sections instantaneous velocity
continuity elementary function 2.1: Instantaneous Rates of Change- The Derivative
5.5: Numerical Integration 3.2: The Mean Value Theorem
1.5: Continuity
12.2: Limits and Continuity of Multivariable Ellipses Integral Test
Functions 9.1: Conic Sections 8.3: Integral and Comparison Tests
contour map extrema Integration by Parts
12.1: Introduction to Multivariable Functions 3.1: Extreme Values 6.2: Integration by Parts
convergence of a series 12.8: Extreme Values Intermediate Value Theorem
8.2: Infinite Series Extreme Value Theorem 1.5: Continuity
cross product 3.1: Extreme Values iterated integration
10.4: The Cross Product 13.1: Iterated Integrals and Area

1 https://math.libretexts.org/@go/page/39399
L orthogonal projection second derivative test
L'Hôpital's Rule 10.3: The Dot Product 3.4: Concavity and the Second Derivative
osculating circle 12.8: Extreme Values
6.7: L'Hopital's Rule
11.5: The Arc Length Parameter and Curvature second partial derivatives
lamina
12.3: Partial Derivatives
13.4: Center of Mass
P sequences
level curve of a function of two variables
8: Sequences and Series
12.1: Introduction to Multivariable Functions parabola 8.1: Sequences
level curves 9.1: Conic Sections
series
12.1: Introduction to Multivariable Functions Parallel Line 8: Sequences and Series
level surface of a function of three 10.5: Lines
Shell Method
variables Parallelogram law 7.3: The Shell Method
10.2: An Introduction to Vectors
12.1: Introduction to Multivariable Functions Simpson's rule
limacon Parametric equations 5.5: Numerical Integration
9.2: Parametric Equations
9.5: Calculus and Polar Functions skew lines
limit parametric equations of a line 10.5: Lines
10.5: Lines
1.2: Epsilon-Delta Definition of a Limit solid of revolution
limits partial derivative 7.4: Arc Length and Surface Area
12.3: Partial Derivatives 9.3: Calculus and Parametric Equations
1.1: An Introduction to Limits
1.3: Finding Limits Analytically partial fraction decomposition speed
Limits of Piecewise functions 6.5: Partial Fraction Decomposition 11.3: The Calculus of Motion
1.4: One Sided Limits partial sum squeeze theorem
local extremum 8.2: Infinite Series 1.3: Finding Limits Analytically
3.1: Extreme Values perimeter Sum Rule
Logarithmic Differentiation 4.3: Optimization 5.1: Antiderivatives and Indefinite Integration
2.6: Implicit Differentiation plane surface area
10.1: Introduction to Cartesian Coordinates in Space 7.4: Arc Length and Surface Area
M 10.6: Planes 9.3: Calculus and Parametric Equations
polar coordinates 9.5: Calculus and Polar Functions
Maclaurin series 13.5: Surface Area
9.4: Introduction to Polar Coordinates
8.8: Taylor Series surface of revolution
power rule
maxima 9.3: Calculus and Parametric Equations
2.3: Basic Differentiation Rules
4.3: Optimization 2.6: Implicit Differentiation 10.1: Introduction to Cartesian Coordinates in Space
mean value theorem 5.1: Antiderivatives and Indefinite Integration
3.2: The Mean Value Theorem power series T
midpoint rule 8.6: Power Series Tangent Line
5.3: Riemann Sums pressure 12.7: Tangent Lines, Normal Lines, and Tangent
Minima 7.6: Fluid Forces Planes
4.3: Optimization projectile motion tangent plane
12.8: Extreme Values 11.3: The Calculus of Motion 12.7: Tangent Lines, Normal Lines, and Tangent
mixed partial derivatives projection Planes
12.3: Partial Derivatives 10.3: The Dot Product tangents (parametric equations)
Multiple Integration 9.3: Calculus and Parametric Equations
13.1: Iterated Integrals and Area Q Taylor polynomials
Multivariable Chain Rule quotient rule 8.7: Taylor Polynomials
12.5: The Multivariable Chain Rule
2.4: The Product and Quotient Rules
Taylor series
Multivariable Functions 8.8: Taylor Series
12.1: Introduction to Multivariable Functions
R telescoping series
8.2: Infinite Series
N ratio test the Product Rule
8.4: Ratio and Root Tests
Newton's Method related rates
2.4: The Product and Quotient Rules
4.1: Newton's Method Total differential
4.2: Related Rates
normal lines (parametric equations) Riemann sums
12.4: Differentiability and the Total Differential
9.3: Calculus and Parametric Equations Trigonometric Integrals
5.3: Riemann Sums
normal vector Rolle’s Theorem
6.3: Trigonometric Integrals
10.6: Planes triple integral
3.2: The Mean Value Theorem
11.4: Unit Tangent and Normal Vectors 13.6: Volume Between Surfaces and Triple
root test Integration
O 8.4: Ratio and Root Tests

one sided limit U


1.4: One Sided Limits
S Unit Tangent Vector
Optimization saddle point 11.4: Unit Tangent and Normal Vectors
12.8: Extreme Values
4.3: Optimization
orthogonal decomposition Second Derivative
3.4: Concavity and the Second Derivative
10.3: The Dot Product

2 https://math.libretexts.org/@go/page/39399
V velocity W
vector 11.3: The Calculus of Motion washer method
10.2: An Introduction to Vectors Volume Between Surfaces 7.2: Volume by Cross-Sectional Area- Disk and
vector function 13.6: Volume Between Surfaces and Triple Washer Methods
Integration work
11.1: Vector–Valued Functions
Volume by Shells 7.5: Work
vectors
7.3: The Shell Method
10: Vectors

3 https://math.libretexts.org/@go/page/39399
Glossary
average rate of change | is a function f(x) over an characteristic equation | the equation
absolute convergence | if the series \displaystyle interval [x,x+h] is \frac{f(x+h)−f(a)}{b−a} aλ^2+bλ+c=0 for the differential equation ay″+by′
\sum^∞_{n=1}|a_n| converges, the series \displaystyle
+cy=0
\sum^∞_{n=1}a_n is said to converge absolutely average value of a function | (or f_{ave}) the
average value of a function on an interval can be found circulation | the tendency of a fluid to move in the
absolute error | if B is an estimate of some by calculating the definite integral of the function and direction of curve C. If C is a closed curve, then the
quantity having an actual value of A, then the absolute
dividing that value by the length of the interval circulation of \vecs F along C is line integral ∫_C \vecs
error is given by |A−B|
F·\vecs T \,ds, which we also denote ∮_C\vecs F·\vecs
average velocity | the change in an object’s T \,ds.
absolute extremum | if f has an absolute position divided by the length of a time period; the
maximum or absolute minimum at c, we say f has an average velocity of an object over a time interval [t,a] closed curve | a curve for which there exists a
absolute extremum at c (if t<a or [a,t] if t>a), with a position given by s(t), that parameterization \vecs r(t), a≤t≤b, such that \vecs
absolute maximum | if f(c)≥f(x) for all x in the is v_{ave}=\dfrac{s(t)−s(a)}{t−a} r(a)=\vecs r(b), and the curve is traversed exactly once
domain of f, we say f has an absolute maximum at c
base | the number b in the exponential function closed curve | a curve that begins and ends at the
absolute minimum | if f(c)≤f(x) for all x in the f(x)=b^x and the logarithmic function f(x)=\log_bx same point
domain of f, we say f has an absolute minimum at c binomial series | the Maclaurin series for f(x)= closed set | a set S that contains all its boundary
absolute value function | f(x)=\begin{cases}−x, (1+x)^r; it is given by points
& \text{if } x<0\x, & \text{if } x≥0\end{cases} (1+x)^r=\sum_{n=0}^∞(^r_n)x^n=1+rx+\dfrac{r(r−1)
}{2!}x^2+⋯+\dfrac{r(r−1)⋯(r−n+1)}{n!}x^n+⋯ for comparison test | If 0≤a_n≤b_n for all n≥N and
acceleration | is the rate of change of the velocity, |x|<1 \displaystyle \sum^∞_{n=1}b_n converges, then
that is, the derivative of velocity \displaystyle \sum^∞_{n=1}a_n converges; if
binormal vector | a unit vector orthogonal to the a_n≥b_n≥0 for all n≥N and \displaystyle
acceleration vector | the second derivative of the unit tangent vector and the unit normal vector \sum^∞_{n=1}b_n diverges, then \displaystyle
position vector \sum^∞_{n=1}a_n diverges.
boundary conditions | the conditions that give the
algebraic function | a function involving any state of a system at different times, such as the position complementary equation | for the
combination of only the basic operations of addition, of a spring-mass system at two different times nonhomogeneous linear differential equation a+2(x)y″
subtraction, multiplication, division, powers, and roots +a_1(x)y′+a_0(x)y=r(x), \nonumber the associated
applied to an input variable x boundary point | a point P_0 of R is a boundary homogeneous equation, called the complementary
point if every δ disk centered around P_0 contains
alternating series | a series of the form equation, is a_2(x)y''+a_1(x)y′+a_0(x)y=0 \nonumber
points both inside and outside R
\displaystyle \sum^∞_{n=1}(−1)^{n+1}b_n or component | a scalar that describes either the
\displaystyle \sum^∞_{n=1}(−1)^nb_n, where b_n≥0, boundary-value problem | a differential vertical or horizontal direction of a vector
is called an alternating series equation with associated boundary conditions
component functions | the component functions
alternating series test | for an alternating series of bounded above | a sequence \displaystyle {a_n} is of the vector-valued function \vecs
either form, if b_{n+1}≤b_n for all integers n≥1 and bounded above if there exists a constant \displaystyle
r(t)=f(t)\hat{\mathbf{i}}+g(t)\hat{\mathbf{j}} are f(t)
b_n→0, then an alternating series converges M such that \displaystyle a_n≤M for all positive
and g(t), and the component functions of the vector-
integers \displaystyle n
amount of change | the amount of a function f(x) valued function \vecs
over an interval [x,x+h] is f(x+h)−f(x) bounded below | a sequence \displaystyle {a_n} is r(t)=f(t)\hat{\mathbf{i}}+g(t)\hat{\mathbf{j}}+h(t)\ha
bounded below if there exists a constant \displaystyle t{\mathbf{k}} are f(t), g(t) and h(t)
angular coordinate | θ the angle formed by a line M such that \displaystyle M≤a_n for all positive
segment connecting the origin to a point in the polar composite function | given two functions f and g,
integers \displaystyle n
coordinate system with the positive radial (x) axis, a new function, denoted g∘f, such that (g∘f)
measured counterclockwise bounded sequence | a sequence \displaystyle (x)=g(f(x))
{a_n} is bounded if there exists a constant computer algebra system (CAS) | technology
antiderivative | a function F such that F′(x)=f(x) \displaystyle M such that \displaystyle |a_n|≤M for all
for all x in the domain of f is an antiderivative of f used to perform many mathematical tasks, including
positive integers \displaystyle n integration
arc length | the arc length of a curve can be thought cardioid | a plane curve traced by a point on the
of as the distance a person would travel along the path concave down | if f is differentiable over an interval
perimeter of a circle that is rolling around a fixed
of the curve I and f' is decreasing over I, then f is concave down
circle of the same radius; the equation of a cardioid is
over I
arc-length function | a function s(t) that describes r=a(1+\sin θ) or r=a(1+\cos θ)
the arc length of curve C as a function of t concave up | if f is differentiable over an interval I
carrying capacity | the maximum population of an and f' is increasing over I, then f is concave up over I
arc-length parameterization a | organism that the environment can sustain indefinitely
reparameterization of a vector-valued function in concavity | the upward or downward curve of the
catenary | a curve in the shape of the function graph of a function
which the parameter is equal to the arc length y=a\cdot\cosh(x/a) is a catenary; a cable of uniform
arithmetic sequence | a sequence in which the density suspended between two supports assumes the concavity test | suppose f is twice differentiable
difference between every pair of consecutive terms is shape of a catenary over an interval I; if f''>0 over I, then f is concave up
the same is called an arithmetic sequence over I; if f''< over I, then f is concave down over I
center of mass | the point at which the total mass of
asymptotically semi-stable solution | y=k if it the system could be concentrated without changing the conditional convergence | if the series
is neither asymptotically stable nor asymptotically moment \displaystyle \sum^∞_{n=1}a_n converges, but the
unstable series \displaystyle \sum^∞_{n=1}|a_n| diverges, the
centroid | the centroid of a region is the geometric series \displaystyle \sum^∞_{n=1}a_n is said to
asymptotically stable solution | y=k if there center of the region; laminas are often represented by converge conditionally
exists ε>0 such that for any value c∈(k−ε,k+ε) the regions in the plane; if the lamina has a constant
solution to the initial-value problem y′=f(x,y),y(x_0)=c density, the center of mass of the lamina depends only conic section | a conic section is any curve formed
approaches k as x approaches infinity on the shape of the corresponding planar region; in this by the intersection of a plane with a cone of two
case, the center of mass of the lamina corresponds to nappes
asymptotically unstable solution | y=k if there the centroid of the representative region
exists ε>0 such that for any value c∈(k−ε,k+ε) the connected region | a region in which any two
solution to the initial-value problem y′=f(x,y),y(x_0)=c chain rule | the chain rule defines the derivative of a points can be connected by a path with a trace
never approaches k as xapproaches infinity composite function as the derivative of the outer contained entirely inside the region
function evaluated at the inner function times the
autonomous differential equation | an derivative of the inner function connected set | an open set S that cannot be
equation in which the right-hand side is a function of y represented as the union of two or more disjoint,
alone change of variables | the substitution of a nonempty open subsets
variable, such as u, for an expression in the integrand
conservative field | a vector field for which there
exists a scalar function f such that \vecs ∇f=\vecs{F}

1 https://math.libretexts.org/@go/page/51385
constant multiple law for limits | the limit law curl | the curl of vector field \vecs{F}=⟨P,Q,R⟩, differentiable | a function f(x,y) is differentiable at
\lim_{x→a}cf(x)=c⋅\lim_{x→a}f(x)=cL \nonumber denoted \vecs ∇× \vecs{F} is the “determinant” of the (x_0,y_0) if f(x,y) can be expressed in the form
matrix \begin{vmatrix} \mathbf{\hat i} & f(x,y)=f(x_0,y_0)+f_x(x_0,y_0)(x−x_0)+f_y(x_0,y_0)
constant multiple rule | the derivative of a \mathbf{\hat j} & \mathbf{\hat k} \ \dfrac{\partial} (y−y_0)+E(x,y), where the error term E(x,y) satisfies
constant c multiplied by a function f is the same as the
{\partial x} & \dfrac{\partial}{\partial y} & \lim_{(x,y)→(x_0,y_0)}\dfrac{E(x,y)}
constant multiplied by the derivative: \dfrac{d}
\dfrac{\partial}{\partial z} \ P & Q & R {\sqrt{(x−x_0)^2+(y−y_0)^2}}=0
{dx}\big(cf(x)\big)=cf′(x)
\end{vmatrix}. \nonumber and is given by the
expression (R_y−Q_z)\,\mathbf{\hat i} +
differentiable at a | a function for which f'(a)
constant rule | the derivative of a constant function exists is differentiable at a
is zero: \dfrac{d}{dx}(c)=0, where c is a constant (P_z−R_x)\,\mathbf{\hat j} +(Q_x−P_y)\,\mathbf{\hat
k} ; it measures the tendency of particles at a point to differentiable function | a function for which
constraint | an inequality or equation involving one rotate about the axis that points in the direction of the f'(x) exists is a differentiable function
or more variables that is used in an optimization curl at the point
problem; the constraint enforces a limit on the possible differentiable on S | a function for which f'(x)
solutions for the problem curvature | the derivative of the unit tangent vector exists for each x in the open set S is differentiable on S
with respect to the arc-length parameter
continuity at a point | A function f(x) is differential | the differential dx is an independent
continuous at a point a if and only if the following cusp | a pointed end or part where two curves meet variable that can be assigned any nonzero real number;
three conditions are satisfied: (1) f(a) is defined, (2) cycloid | the curve traced by a point on the rim of a the differential dy is defined to be dy=f'(x)\,dx
\displaystyle \lim_{x→a}f(x) exists, and (3) circular wheel as the wheel rolls along a straight line
\displaystyle \lim{x→a}f(x)=f(a)
differential calculus | the field of calculus
without slippage concerned with the study of derivatives and their
continuity from the left | A function is cylinder | a set of lines parallel to a given line applications
continuous from the left at b if \displaystyle
\lim_{x→b^−}f(x)=f(b)
passing through a given curve differential equation | an equation involving a
function y=y(x) and one or more of its derivatives
cylindrical coordinate system | a way to
continuity from the right | A function is describe a location in space with an ordered triple
continuous from the right at a if \displaystyle differential form | given a differentiable function
(r,θ,z), where (r,θ) represents the polar coordinates of y=f'(x), the equation dy=f'(x)\,dx is the differential
\lim_{x→a^+}f(x)=f(a)
the point’s projection in the xy-plane, and z represents form of the derivative of y with respect to x
continuity over an interval | a function that can the point’s projection onto the z-axis
be traced with a pencil without lifting the pencil; a
differentiation | the process of taking a derivative
decreasing on the interval I | a function
function is continuous over an open interval if it is decreasing on the interval I if, for all direction angles | the angles formed by a nonzero
continuous at every point in the interval; a function x_1,\,x_2∈I,\;f(x_1)≥f(x_2) if x_1<x_2 vector and the coordinate axes
f(x) is continuous over a closed interval of the form
[a,b] if it is continuous at every point in (a,b), and it is definite integral | a primary operation of calculus; direction cosines | the cosines of the angles formed
continuous from the right at a and from the left at b the area between the curve and the x-axis over a given by a nonzero vector and the coordinate axes
interval is a definite integral direction field (slope field) | a mathematical
contour map | a plot of the various level curves of a
given function f(x,y) definite integral of a vector-valued function object used to graphically represent solutions to a first-
| the vector obtained by calculating the definite order differential equation; at each point in a direction
convergence of a series | a series converges if the integral of each of the component functions of a given field, a line segment appears whose slope is equal to
sequence of partial sums for that series converges vector-valued function, then using the results as the the slope of a solution to the differential equation
components of the resulting function passing through that point
convergent sequence | a convergent sequence is a
sequence \displaystyle {a_n} for which there exists a degree | for a polynomial function, the value of the direction vector | a vector parallel to a line that is
real number \displaystyle L such that \displaystyle a_n largest exponent of any term used to describe the direction, or orientation, of the
is arbitrarily close to \displaystyle L as long as line in space
\displaystyle n is sufficiently large density function | a density function describes how
mass is distributed throughout an object; it can be a directional derivative | the derivative of a
coordinate plane | a plane containing two of the linear density, expressed in terms of mass per unit function in the direction of a given unit vector
three coordinate axes in the three-dimensional length; an area density, expressed in terms of mass per
coordinate system, named by the axes it contains: the directrix | a directrix (plural: directrices) is a line
unit area; or a volume density, expressed in terms of used to construct and define a conic section; a parabola
xy-plane, xz-plane, or the yz-plane mass per unit volume; weight-density is also used to has one directrix; ellipses and hyperbolas have two
critical point | if f'(c)=0 or f'(c) is undefined, we describe weight (rather than mass) per unit volume
say that c is a critical point of f discontinuity at a point | A function is
dependent variable | the output variable for a discontinuous at a point or has a discontinuity at a
critical point of a function of two variables | function
point if it is not continuous at the point
the point (x_0,y_0) is called a critical point of f(x,y) if
derivative | the slope of the tangent line to a discriminant | the value 4AC−B^2, which is used
one of the two following conditions holds: 1.
function at a point, calculated by taking the limit of the to identify a conic when the equation contains a term
f_x(x_0,y_0)=f_y(x_0,y_0)=0 2. At least one of
difference quotient, is the derivative involving xy, is called a discriminant
f_x(x_0,y_0) and f_y(x_0,y_0) do not exist
derivative function | gives the derivative of a discriminant | the discriminant of the function
cross product | \vecs u×\vecs v=
function at each point in the domain of the original
(u_2v_3−u_3v_2)\mathbf{\hat i}− f(x,y) is given by the formula D=f_{xx}
function for which the derivative is defined
(u_1v_3−u_3v_1)\mathbf{\hat j}+ (x_0,y_0)f_{yy}(x_0,y_0)−(f_{xy}(x_0,y_0))^2
(u_1v_2−u_2v_1)\mathbf{\hat k}, where \vecs derivative of a vector-valued function | the disk method | a special case of the slicing method
u=⟨u_1,u_2,u_3⟩ and \vecs v=⟨v_1,v_2,v_3⟩ derivative of a vector-valued function \vecs{r}(t) is
used with solids of revolution when the slices are disks
determinant a real number associated with a square \vecs{r}′(t) = \lim \limits_{\Delta t \to 0} \frac{\vecs
matrix parallelepiped a three-dimensional prism with r(t+\Delta t)−\vecs r(t)}{ \Delta t}, provided the limit divergence | the divergence of a vector field
six faces that are parallelograms torque the effect of a exists \vecs{F}=⟨P,Q,R⟩, denoted \vecs ∇× \vecs{F}, is
force that causes an object to rotate triple scalar P_x+Q_y+R_z; it measures the “outflowing-ness” of a
product the dot product of a vector with the cross difference law for limits | the limit law vector field
product of two other vectors: \vecs u⋅(\vecs v×\vecs w) \lim_{x→a}(f(x)−g(x))=\lim_{x→a}f(x)−
vector product the cross product of two vectors. \lim_{x→a}g(x)=L−M \nonumber divergence of a series | a series diverges if the
sequence of partial sums for that series diverges
cross-section | the intersection of a plane and a solid difference quotient | of a function f(x) at a is
object
given by \dfrac{f(a+h)−f(a)}{h} or \dfrac{f(x)−f(a)} divergence test | if \displaystyle
{x−a} \lim_{n→∞}a_n≠0, then the series \displaystyle
cubic function | a polynomial of degree 3; that is, a \sum^∞_{n=1}a_n diverges
function of the form f(x)=ax^3+bx^2+cx+d, where a≠0 difference rule | the derivative of the difference of
a function f and a function g is the same as the divergent sequence | a sequence that is not
difference of the derivative of f and the derivative of g: convergent is divergent
\dfrac{d}{dx}\big(f(x)−g(x)\big)=f′(x)−g′(x)
domain | the set of inputs for a function

2 https://math.libretexts.org/@go/page/51385
dot product or scalar product | \vecs{ u} flux | the rate of a fluid flowing across a curve in a geometric series | a geometric series is a series that
⋅\vecs{ v}=u_1v_1+u_2v_2+u_3v_3 where \vecs{ vector field; the flux of vector field \vecs F across can be written in the form \displaystyle
u}=⟨u_1,u_2,u_3⟩ and \vecs{ v}=⟨v_1,v_2,v_3⟩ plane curve C is line integral ∫_C \vecs F·\frac{\vecs \sum_{n=1}^∞ar^{n−1}=a+ar+ar^2+ar^3+⋯
n(t)}{‖\vecs n(t)‖} \,ds
double integral | of the function f(x,y) over the gradient | the gradient of the function f(x,y) is
region R in the xy-plane is defined as the limit of a flux integral | another name for a surface integral of defined to be \vecs ∇f(x,y)=(∂f/∂x)\,\hat{\mathbf i}+
double Riemann sum, \iint_R f(x,y) \,dA = a vector field; the preferred term in physics and (∂f/∂y)\,\hat{\mathbf j}, which can be generalized to a
\lim_{m,n\rightarrow \infty} \sum_{i=1}^m engineering function of any number of independent variables
\sum_{j=1}^n f(x_{ij}^*, y_{ij}^*) \,\Delta A.
\nonumber focal parameter | the focal parameter is the gradient field | a vector field \vecs{F} for which
distance from a focus of a conic section to the nearest there exists a scalar function f such that \vecs
double Riemann sum | of the function f(x,y) over directrix ∇f=\vecs{F}; in other words, a vector field that is the
a rectangular region R is \sum_{i=1}^m \sum_{j=1}^n gradient of a function; such vector fields are also
f(x_{ij}^*, y_{ij}^*) \,\Delta A, \nonumber where R is
focus | a focus (plural: foci) is a point used to called conservative
construct and define a conic section; a parabola has
divided into smaller subrectangles R_{ij} and
(x_{ij}^*, y_{ij}^*) is an arbitrary point in R_{ij}
one focus; an ellipse and a hyperbola have two graph of a function | the set of points (x,y) such
that x is in the domain of f and y=f(x)
doubling time | if a quantity grows exponentially, formal definition of an infinite limit |
the doubling time is the amount of time it takes the
\displaystyle \lim_{x→a}f(x)=\infty if for every M>0, graph of a function of two variables | a set of
there exists a δ>0 such that if 0<|x−a|<δ, then f(x)>M ordered triples (x,y,z) that satisfies the equation
quantity to double, and is given by (\ln 2)/k
\displaystyle \lim_{x→a}f(x)=-\infty if for every M>0, z=f(x,y) plotted in three-dimensional Cartesian space
eccentricity | the eccentricity is defined as the there exists a δ>0 such that if 0<|x−a|<δ, then f(x)<-M
distance from any point on the conic section to its
Green’s theorem | relates the integral over a
Frenet frame of reference | (TNB frame) a connected region to an integral over the boundary of
focus divided by the perpendicular distance from that
frame of reference in three-dimensional space formed the region
point to the nearest directrix
by the unit tangent vector, the unit normal vector, and
ellipsoid | a three-dimensional surface described by the binormal vector grid curves | curves on a surface that are parallel to
grid lines in a coordinate plane
an equation of the form \dfrac{x^2}{a^2}+\dfrac{y^2}
frustum | a portion of a cone; a frustum is
{b^2}+\dfrac{z^2}{c^2}=1; all traces of this surface growth rate | the constant r>0 in the exponential
constructed by cutting the cone with a plane parallel to
are ellipses growth function P(t)=P_0e^{rt}
the base
elliptic cone | a three-dimensional surface described half-life | if a quantity decays exponentially, the half-
by an equation of the form \dfrac{x^2}
Fubini’s theorem | if f(x,y) is a function of two
variables that is continuous over a rectangular region R life is the amount of time it takes the quantity to be
{a^2}+\dfrac{y^2}{b^2}−\dfrac{z^2}{c^2}=0; traces reduced by half. It is given by (\ln 2)/k
= \big\{(x,y) \in \mathbb{R}^2 \,|\,a \leq x \leq b, \, c
of this surface include ellipses and intersecting lines
\leq y \leq d\big\}, then the double integral of f over harmonic series | the harmonic series takes the
elliptic paraboloid | a three-dimensional surface the region equals an iterated integral, form \displaystyle \sum_{n=1}^∞\frac{1}
described by an equation of the form z=\dfrac{x^2} \displaystyle\iint_R f(x,y) \, dA = \int_a^b \int_c^d {n}=1+\frac{1}{2}+\frac{1}{3}+⋯
{a^2}+\dfrac{y^2}{b^2}; traces of this surface include f(x,y) \,dx \, dy = \int_c^d \int_a^b f(x,y) \,dx \, dy
ellipses and parabolas \nonumber heat flow | a vector field proportional to the negative
temperature gradient in an object
end behavior | the behavior of a function as x→∞ function | a set of inputs, a set of outputs, and a rule
and x→−∞ for mapping each input to exactly one output helix | a three-dimensional curve in the shape of a
spiral
epsilon-delta definition of the limit | function of two variables | a function z=f(x,y)
\displaystyle \lim_{x→a}f(x)=L if for every ε>0, there that maps each ordered pair (x,y) in a subset D of R^2 higher-order derivative | a derivative of a
exists a δ>0 such that if 0<|x−a|<δ, then |f(x)−L|<ε to a unique real number z derivative, from the second derivative to the
n^{\text{th}} derivative, is called a higher-order
equilibrium solution | any solution to the Fundamental Theorem for Line Integrals | derivative
differential equation of the form y=c, where c is a the value of line integral \displaystyle \int_C\vecs
constant ∇f⋅d\vecs r depends only on the value of f at the higher-order partial derivatives | second-order
endpoints of C: \displaystyle \int_C \vecs ∇f⋅d\vecs or higher partial derivatives, regardless of whether
equivalent vectors | vectors that have the same r=f(\vecs r(b))−f(\vecs r(a)) they are mixed partial derivatives
magnitude and the same direction
fundamental theorem of calculus | (also, homogeneous linear equation | a second-order
Euler’s Method | a numerical technique used to evaluation theorem) we can evaluate a definite integral differential equation that can be written in the form
approximate solutions to an initial-value problem by evaluating the antiderivative of the integrand at the a_2(x)y″+a_1(x)y′+a_0(x)y=r(x), but r(x)=0 for every
even function | a function is even if f(−x)=f(x) for endpoints of the interval and subtracting value of x
all x in the domain of f fundamental theorem of calculus | uses a Hooke’s law | this law states that the force required
explicit formula | a sequence may be defined by an definite integral to define an antiderivative of a to compress (or elongate) a spring is proportional to
explicit formula such that \displaystyle a_n=f(n) function the distance the spring has been compressed (or
stretched) from equilibrium; in other words, F=kx,
exponent | the value x in the expression b^x fundamental theorem of calculus | the where k is a constant
theorem, central to the entire development of calculus,
exponential decay | systems that exhibit that establishes the relationship between differentiation horizontal asymptote | if \displaystyle
exponential decay follow a model of the form and integration \lim_{x→∞}f(x)=L or \displaystyle
y=y_0e^{−kt} \lim_{x→−∞}f(x)=L, then y=L is a horizontal
general form | an equation of a conic section asymptote of f
exponential growth | systems that exhibit written as a general second-degree equation
exponential growth follow a model of the form horizontal line test | a function f is one-to-one if
y=y_0e^{kt} general form of the equation of a plane | an and only if every horizontal line intersects the graph of
equation in the form ax+by+cz+d=0, where \vecs f, at most, once
extreme value theorem | if f is a continuous n=⟨a,b,c⟩ is a normal vector of the plane, P=
function over a finite, closed interval, then f has an (x_0,y_0,z_0) is a point on the plane, and hydrostatic pressure | the pressure exerted by
absolute maximum and an absolute minimum d=−ax_0−by_0−cz_0 water on a submerged object
Fermat’s theorem | if f has a local extremum at c, general solution (or family of solutions) | the hyperbolic functions | the functions denoted
then c is a critical point of f entire set of solutions to a given differential equation \sinh,\,\cosh,\,\operatorname{tanh},\,\operatorname{cs
ch},\,\operatorname{sech}, and \coth, which involve
first derivative test | let f be a continuous function generalized chain rule | the chain rule extended certain combinations of e^x and e^{−x}
over an interval I containing a critical point c such that to functions of more than one independent variable, in
f is differentiable over I except possibly at c; if f' which each independent variable may depend on one hyperboloid of one sheet | a three-dimensional
changes sign from positive to negative as x increases or more other variables surface described by an equation of the form
through c, then f has a local maximum at c; if f' \dfrac{x^2}{a^2}+\dfrac{y^2}{b^2}−\dfrac{z^2}
changes sign from negative to positive as x increases geometric sequence | a sequence \displaystyle {c^2}=1; traces of this surface include ellipses and
through c, then f has a local minimum at c; if f' does {a_n} in which the ratio \displaystyle a_{n+1}/a_n is hyperbolas
not change sign as x increases through c, then f does the same for all positive integers \displaystyle n is
not have a local extremum at c called a geometric sequence

3 https://math.libretexts.org/@go/page/51385
hyperboloid of two sheets | a three-dimensional initial-value problem | a differential equation iterated integral | for a function f(x,y) over the
surface described by an equation of the form together with an initial value or values region R is a. \displaystyle \int_a^b \int_c^d f(x,y) \,dx
\dfrac{z^2}{c^2}−\dfrac{x^2}{a^2}−\dfrac{y^2} \, dy = \int_a^b \left[\int_c^d f(x,y) \, dy\right] \, dx, b.
{b^2}=1; traces of this surface include ellipses and
instantaneous rate of change | the rate of \displaystyle \int_c^d \int_a^b f(x,y) \, dx \, dy =
change of a function at any point along the function a,
hyperbolas \int_c^d \left[\int_a^b f(x,y) \, dx\right] \, dy, where
also called f′(a), or the derivative of the function at a
a,b,c, and d are any real numbers and R = [a,b] \times
implicit differentiation | is a technique for [c,d]
computing \dfrac{dy}{dx} for a function defined by instantaneous velocity | The instantaneous
an equation, accomplished by differentiating both sides velocity of an object with a position function that is iterative process | process in which a list of
of the equation (remembering to treat the variable y as given by s(t) is the value that the average velocities on numbers x_0,x_1,x_2,x_3… is generated by starting
a function) and solving for \dfrac{dy}{dx} intervals of the form [t,a] and [a,t] approach as the with a number x_0 and defining x_n=F(x_{n−1}) for
values of t move closer to a, provided such a value n≥1
improper double integral | a double integral exists
over an unbounded region or of an unbounded function Jacobian | the Jacobian J (u,v) in two variables is a 2
integrable function | a function is integrable if the \times 2 determinant: J(u,v) = \begin{vmatrix}
improper integral | an integral over an infinite limit defining the integral exists; in other words, if the
\frac{\partial x}{\partial u} \frac{\partial y}{\partial u}
interval or an integral of a function containing an limit of the Riemann sums as n goes to infinity exists
\nonumber \ \frac{\partial x}{\partial v} \frac{\partial
infinite discontinuity on the interval; an improper
integral calculus | the study of integrals and their y}{\partial v} \end{vmatrix}; \nonumber the Jacobian
integral is defined in terms of a limit. The improper
applications J (u,v,w) in three variables is a 3 \times 3 determinant:
integral converges if this limit is a finite real number;
J(u,v,w) = \begin{vmatrix} \frac{\partial x}{\partial u}
otherwise, the improper integral diverges integral test | for a series \displaystyle \frac{\partial y}{\partial u} \frac{\partial z}{\partial u}
\sum^∞_{n=1}a_n with positive terms a_n, if there \nonumber \ \frac{\partial x}{\partial v} \frac{\partial
increasing on the interval I | a function
exists a continuous, decreasing function f such that y}{\partial v} \frac{\partial z}{\partial v} \nonumber \
increasing on the interval I if for all
f(n)=a_n for all positive integers n, then \frac{\partial x}{\partial w} \frac{\partial y}{\partial
x_1,\,x_2∈I,\;f(x_1)≤f(x_2) if x_1<x_2
\sum_{n=1}^∞a_n \nonumber and ∫^∞_1f(x)\,dx w} \frac{\partial z}{\partial w}\end{vmatrix}
indefinite integral | the most general \nonumber either both converge or both diverge \nonumber
antiderivative of f(x) is the indefinite integral of f; we
use the notation \displaystyle \int f(x)\,dx to denote the
integrand | the function to the right of the jump discontinuity | A jump discontinuity occurs
integration symbol; the integrand includes the function at a point a if \displaystyle \lim_{x→a^−}f(x) and
indefinite integral of f
being integrated \displaystyle \lim_{x→a^+}f(x) both exist, but
indefinite integral of a vector-valued \displaystyle \lim_{x→a^−}f(x)≠\lim_{x→a^+}f(x)
function | a vector-valued function with a derivative integrating factor | any function f(x) that is
that is equal to a given vector-valued function multiplied on both sides of a differential equation to Kepler’s laws of planetary motion | three laws
make the side involving the unknown function equal to governing the motion of planets, asteroids, and comets
independence of path | a vector field \vecs{F} the derivative of a product of two functions in orbit around the Sun
has path independence if \displaystyle \int_{C_1}
\vecs F⋅d\vecs r=\displaystyle \int_{C_2} \vecs integration by parts | a technique of integration Lagrange multiplier | the constant (or constants)
F⋅d\vecs r for any curves C_1 and C_2 in the domain that allows the exchange of one integral for another used in the method of Lagrange multipliers; in the case
of \vecs{F} with the same initial points and terminal using the formula \displaystyle ∫u\,dv=uv−∫v\,du of one constant, it is represented by the variable λ
points integration by substitution | a technique for lamina | a thin sheet of material; laminas are thin
independent variable | the input variable for a integration that allows integration of functions that are enough that, for mathematical purposes, they can be
function the result of a chain-rule derivative treated as if they are two-dimensional
indeterminate forms | When evaluating a limit, integration table | a table that lists integration left-endpoint approximation | an approximation
the forms \dfrac{0}{0},∞/∞, 0⋅∞, ∞−∞, 0^0, ∞^0, and formulas of the area under a curve computed by using the left
1^∞ are considered indeterminate because further interior point | a point P_0 of \mathbb{R} is a endpoint of each subinterval to calculate the height of
analysis is required to determine whether the limit boundary point if there is a δ disk centered around P_0 the vertical sides of each rectangle
exists and, if so, what its value is. contained completely in \mathbb{R} level curve of a function of two variables |
index variable | the subscript used to define the the set of points satisfying the equation f(x,y)=c for
Intermediate Value Theorem | Let f be
terms in a sequence is called the index some real number c in the range of f
continuous over a closed bounded interval [a,b] if z is
infinite discontinuity | An infinite discontinuity any real number between f(a) and f(b), then there is a level surface of a function of three variables
occurs at a point a if \displaystyle number c in [a,b] satisfying f(c)=z | the set of points satisfying the equation f(x,y,z)=c for
\lim_{x→a^−}f(x)=±∞ or \displaystyle some real number c in the range of f
intermediate variable | given a composition of
\lim_{x→a^+}f(x)=±∞ functions (e.g., \displaystyle f(x(t),y(t))), the limaçon | the graph of the equation r=a+b\sin θ or
infinite limit | A function has an infinite limit at a intermediate variables are the variables that are r=a+b\cos θ. If a=b then the graph is a cardioid
point a if it either increases or decreases without bound independent in the outer function but dependent on
other variables as well; in the function \displaystyle limit | the process of letting x or t approach a in an
as it approaches a expression; the limit of a function f(x) as x approaches
f(x(t),y(t)), the variables \displaystyle x and
infinite limit at infinity | a function that becomes \displaystyle y are examples of intermediate variables a is the value that f(x) approaches as x approaches a
arbitrarily large as x becomes large limit at infinity | a function that approaches a limit
interval of convergence | the set of real numbers
infinite series | an infinite series is an expression of x for which a power series converges value L as x becomes large
the form \displaystyle a_1+a_2+a_3+⋯ limit comparison test | Suppose a_n,b_n≥0 for all
=\sum_{n=1}^∞a_n intuitive definition of the limit | If all values of
the function f(x) approach the real number L as the n≥1. If \displaystyle \lim_{n→∞}a_n/b_n→L≠0, then
inflection point | if f is continuous at c and f values of x(≠a) approach a, f(x) approaches L \displaystyle \sum^∞_{n=1}a_n and \displaystyle
changes concavity at c, the point (c,f(c)) is an \sum^∞_{n=1}b_n both converge or both diverge; if
inflection point of f inverse function | for a function f, the inverse \displaystyle \lim_{n→∞}a_n/b_n→0 and
function f^{−1} satisfies f^{−1}(y)=x if f(x)=y \displaystyle \sum^∞_{n=1}b_n converges, then
initial point | the starting point of a vector \displaystyle \sum^∞_{n=1}a_n converges. If
inverse hyperbolic functions | the inverses of \displaystyle \lim_{n→∞}a_n/b_n→∞, and
initial population | the population at time t=0 the hyperbolic functions where \cosh and
\displaystyle \sum^∞_{n=1}b_n diverges, then
\operatorname{sech} are restricted to the domain
initial value problem | a problem that requires \displaystyle \sum^∞_{n=1}a_n diverges.
[0,∞);each of these functions can be expressed in terms
finding a function y that satisfies the differential
equation \dfrac{dy}{dx}=f(x) together with the initial
of a composition of the natural logarithm function and limit laws | the individual properties of limits; for
an algebraic function each of the individual laws, let f(x) and g(x) be defined
condition y(x_0)=y_0
for all x≠a over some open interval containing a;
initial value(s) | a value or set of values that a inverse trigonometric functions | the inverses assume that L and M are real numbers so that
of the trigonometric functions are defined on restricted
solution of a differential equation satisfies for a fixed \lim_{x→a}f(x)=L and \lim_{x→a}g(x)=M; let c be a
domains where they are one-to-one functions
value of the independent variable constant
initial velocity | the velocity at time t=0 limit of a sequence | the real number LL to which
a sequence converges is called the limit of the
sequence

4 https://math.libretexts.org/@go/page/51385
limit of a vector-valued function | a vector- major axis | the major axis of a conic section passes natural logarithm | the function \ln x=\log_ex
valued function \vecs r(t) has a limit \vecs L as t through the vertex in the case of a parabola or through
approaches a if \lim \limits{t \to a} \left| \vecs r(t) - the two vertices in the case of an ellipse or hyperbola;
net change theorem | if we know the rate of
change of a quantity, the net change theorem says the
\vecs L \right| = 0 it is also an axis of symmetry of the conic; also called
future quantity is equal to the initial quantity plus the
the transverse axis
limits of integration | these values appear near the integral of the rate of change of the quantity
top and bottom of the integral sign and define the marginal cost | is the derivative of the cost
interval over which the function should be integrated function, or the approximate cost of producing one net signed area | the area between a function and
more item the x-axis such that the area below the x-axis is
line integral | the integral of a function along a subtracted from the area above the x-axis; the result is
curve in a plane or in space marginal profit | is the derivative of the profit the same as the definite integral of the function
function, or the approximate profit obtained by
linear | description of a first-order differential producing and selling one more item
Newton’s method | method for approximating
equation that can be written in the form a(x)y′ roots of f(x)=0; using an initial guess x_0; each
+b(x)y=c(x) marginal revenue | is the derivative of the revenue subsequent approximation is defined by the equation
function, or the approximate revenue obtained by x_n=x_{n−1}−\frac{f(x_{n−1})}{f'(x_{n−1})}
linear approximation | the linear function selling one more item
L(x)=f(a)+f'(a)(x−a) is the linear approximation of f at nonelementary integral | an integral for which
x=a mass flux | the rate of mass flow of a fluid per unit the antiderivative of the integrand cannot be expressed
area, measured in mass per unit time per unit area as an elementary function
linear approximation | given a function f(x,y)
and a tangent plane to the function at a point mathematical model | A method of simulating nonhomogeneous linear equation | a second-
(x_0,y_0), we can approximate f(x,y) for points near real-life situations with mathematical equations order differential equation that can be written in the
(x_0,y_0) using the tangent plane formula form a_2(x)y″+a_1(x)y′+a_0(x)y=r(x), but r(x)≠0 for
mean value theorem | if f is continuous over [a,b] some value of x
linear function | a function that can be written in and differentiable over (a,b), then there exists c∈(a,b)
the form f(x)=mx+b such that f′(c)=\frac{f(b)−f(a)}{b−a} normal component of acceleration | the
coefficient of the unit normal vector \vecs N when the
linearly dependent | a set of functions mean value theorem for integrals | guarantees acceleration vector is written as a linear combination
f_1(x),f_2(x),…,f_n(x) for whichthere are constants that a point c exists such that f(c) is equal to the of \vecs T and \vecs N
c_1,c_2,…c_n, not all zero, such that average value of the function
c_1f_1(x)+c_2f_2(x)+⋯+c_nf_n(x)=0 for all \(x\) in normal plane | a plane that is perpendicular to a
the interval of interest
method of cylindrical shells | a method of curve at any point on the curve
calculating the volume of a solid of revolution by
linearly independent | a set of functions dividing the solid into nested cylindrical shells; this normal vector | a vector perpendicular to a plane
f_1(x),f_2(x),…,f_n(x) for which there are no method is different from the methods of disks or
constants c_1,c_2,…c_n, such that washers in that we integrate with respect to the normalization | using scalar multiplication to find a
c_1f_1(x)+c_2f_2(x)+⋯+c_nf_n(x)=0 for all \(x\) in opposite variable unit vector with a given direction
the interval of interest number e | as m gets larger, the quantity (1+
method of Lagrange multipliers | a method of
(1/m)^m gets closer to some real number; we define
local extremum | if f has a local maximum or local solving an optimization problem subject to one or
that real number to be e; the value of e is
minimum at c, we say f has a local extremum at c more constraints
approximately 2.718282
local maximum | if there exists an interval I such method of undetermined coefficients | a
that f(c)≥f(x) for all x∈I, we say f has a local method that involves making a guess about the form of
numerical integration | the variety of numerical
methods used to estimate the value of a definite
maximum at c the particular solution, then solving for the coefficients
integral, including the midpoint rule, trapezoidal rule,
in the guess
local minimum | if there exists an interval I such and Simpson’s rule
that f(c)≤f(x) for all x∈I, we say f has a local method of variation of parameters | a method
minimum at c that involves looking for particular solutions in the objective function | the function that is to be
form y_p(x)=u(x)y_1(x)+v(x)y_2(x), where y_1 and maximized or minimized in an optimization problem
logarithmic differentiation | is a technique that y_2 are linearly independent solutions to the
allows us to differentiate a function by first taking the
oblique asymptote | the line y=mx+b if f(x)
complementary equations, and then solving a system approaches it as x→∞ or x→−∞
natural logarithm of both sides of an equation, of equations to find u(x) and v(x)
applying properties of logarithms to simplify the octants | the eight regions of space created by the
equation, and differentiating implicitly midpoint rule | a rule that uses a Riemann sum of coordinate planes
the form \displaystyle M_n=\sum^n_{i=1}f(m_i)Δx,
logarithmic function | a function of the form where m_i is the midpoint of the i^{\text{th}} odd function | a function is odd if f(−x)=−f(x) for
f(x)=\log_b(x) for some base b>0,\,b≠1 such that all x in the domain of f
subinterval to approximate \displaystyle ∫^b_af(x)\,dx
y=\log_b(x) if and only if b^y=x
minor axis | the minor axis is perpendicular to the one-sided limit | A one-sided limit of a function is
logistic differential equation | a differential major axis and intersects the major axis at the center of a limit taken from either the left or the right
equation that incorporates the carrying capacity K and
the conic, or at the vertex in the case of the parabola; one-to-one function | a function f is one-to-one if
growth rate rr into a population model
also called the conjugate axis f(x_1)≠f(x_2) if x_1≠x_2
lower sum | a sum obtained by using the minimum mixed partial derivatives | second-order or
value of f(x) on each subinterval one-to-one transformation | a transformation T :
higher partial derivatives, in which at least two of the G \rightarrow R defined as T(u,v) = (x,y) is said to be
L’Hôpital’s rule | If f and g are differentiable differentiations are with respect to different variables one-to-one if no two points map to the same image
functions over an interval a, except possibly at a, and point
moment | if n masses are arranged on a number line,
\displaystyle \lim_{x→a}f(x)=0=\lim_{x→a}g(x) or
the moment of the system with respect to the origin is open set | a set S that contains none of its boundary
\displaystyle \lim_{x→a}f(x) and \displaystyle
given by \displaystyle M=\sum^n_{i=1}m_ix_i; if, points
\lim_{x→a}g(x) are infinite, then \displaystyle
instead, we consider a region in the plane, bounded
\lim_{x→a}\dfrac{f(x)}{g(x)}=\lim_{x→a}\dfrac{f′ optimization problem | calculation of a
above by a function f(x) over an interval [a,b], then the
(x)}{g′(x)}, assuming the limit on the right exists or is maximum or minimum value of a function of several
moments of the region with respect to the x- and y-
∞ or −∞. variables, often using Lagrange multipliers
axes are given by \displaystyle
Maclaurin polynomial | a Taylor polynomial M_x=ρ∫^b_a\dfrac{[f(x)]^2}{2}\,dx and \displaystyle
optimization problems | problems that are solved
centered at 0; the n^{\text{th}}-degree Taylor M_y=ρ∫^b_axf(x)\,dx, respectively
by finding the maximum or minimum value of a
polynomial for f at 0 is the n^{\text{th}}-degree monotone sequence | an increasing or decreasing function
Maclaurin polynomial for f sequence
order of a differential equation | the highest
Maclaurin series | a Taylor series for a function f multivariable calculus | the study of the calculus order of any derivative of the unknown function that
at x=0 is known as a Maclaurin series for f appears in the equation
of functions of two or more variables
magnitude | the length of a vector nappe | a nappe is one half of a double cone orientation | the direction that a point moves on a
graph as the parameter increases
natural exponential function | the function
f(x)=e^x

5 https://math.libretexts.org/@go/page/51385
orientation of a curve | the orientation of a curve periodic function | a function is periodic if it has a product rule | the derivative of a product of two
C is a specified direction of C repeating pattern as the values of x move from left to functions is the derivative of the first function times
right the second function plus the derivative of the second
orientation of a surface | if a surface has an function times the first function: \dfrac{d}
“inner” side and an “outer” side, then an orientation is phase line | a visual representation of the behavior {dx}\big(f(x)g(x)\big)=f′(x)g(x)+g′(x)f(x)
a choice of the inner or the outer side; the surface of solutions to an autonomous differential equation
could also have “upward” and “downward” subject to various initial conditions projectile motion | motion of an object with an
orientations initial velocity but no force acting on it other than
piecewise smooth curve | an oriented curve that gravity
orthogonal vectors | vectors that form a right is not smooth, but can be written as the union of
angle when placed in standard position finitely many smooth curves propagated error | the error that results in a
calculated quantity f(x) resulting from a measurement
osculating circle | a circle that is tangent to a curve piecewise-defined function | a function that is error dx
C at a point P and that shares the same curvature defined differently on different parts of its domain
quadratic function | a polynomial of degree 2;
osculating plane | the plane determined by the unit planar transformation | a function T that that is, a function of the form f(x)=ax^2+bx+c where
tangent and the unit normal vector transforms a region G in one plane into a region R in a≠0
another plane by a change of variables
p-series | a series of the form \displaystyle quadric surfaces | surfaces in three dimensions
\sum^∞_{n=1}1/n^p plane curve | the set of ordered pairs (f(t),g(t)) having the property that the traces of the surface are
together with their defining parametric equations
parallelogram method | a method for finding the x=f(t) and y=g(t)
conic sections (ellipses, hyperbolas, and parabolas)
sum of two vectors; position the vectors so they share
quotient law for limits | the limit law
the same initial point; the vectors then form two point-slope equation | equation of a linear \lim_{x→a}\dfrac{f(x)}
adjacent sides of a parallelogram; the sum of the function indicating its slope and a point on the graph
{g(x)}=\dfrac{\lim_{x→a}f(x)}
vectors is the diagonal of that parallelogram of the function
{\lim_{x→a}g(x)}=\dfrac{L}{M} for M≠0
parameter | an independent variable that both x and polar axis | the horizontal axis in the polar quotient rule | the derivative of the quotient of two
y depend on in a parametric curve; usually represented coordinate system corresponding to r≥0 functions is the derivative of the first function times
by the variable t the second function minus the derivative of the second
polar coordinate system | a system for locating
parameter domain (parameter space) | the points in the plane. The coordinates are r, the radial function times the first function, all divided by the
region of the uv-plane over which the parameters u and coordinate, and θ, the angular coordinate square of the second function: \dfrac{d}
v vary for parameterization \vecs r(u,v) = \langle {dx}\left(\dfrac{f(x)}{g(x)}\right)=\dfrac{f′(x)g(x)−g′
x(u,v), \, y(u,v), \, z(u,v)\rangle polar equation | an equation or function relating (x)f(x)}{\big(g(x)\big)^2}
the radial coordinate to the angular coordinate in the
parameterization of a curve | rewriting the polar coordinate system radial coordinate | r the coordinate in the polar
equation of a curve defined by a function y=f(x) as coordinate system that measures the distance from a
parametric equations
polar rectangle | the region enclosed between the point in the plane to the pole
circles r = a and r = b and the angles \theta = \alpha
parameterized surface (parametric surface) and \theta = \beta; it is described as R = \{(r, radial field | a vector field in which all vectors
| a surface given by a description of the form \vecs \theta)\,|\,a \leq r \leq b, \, \alpha \leq \theta \leq \beta\} either point directly toward or directly away from the
r(u,v) = \langle x(u,v), \, y(u,v), \, z(u,v)\rangle, where origin; the magnitude of any vector depends only on its
the parameters u and v vary over a parameter domain pole | the central point of the polar coordinate system, distance from the origin
in the uv-plane equivalent to the origin of a Cartesian system
radians | for a circular arc of length s on a circle of
parametric curve | the graph of the parametric polynomial function | a function of the form radius 1, the radian measure of the associated angle θ
equations x(t) and y(t) over an interval a≤t≤b f(x)=a_nx^n+a_{n−1}x^{n−1}+…+a_1x+a_0 is s
combined with the equations population growth rate | is the derivative of the radius of convergence | if there exists a real
parametric equations | the equations x=x(t) and population with respect to time number R>0 such that a power series centered at x=a
y=y(t) that define a parametric curve potential function | a scalar function f such that converges for |x−a|<R and diverges for |x−a|>R, then R
\vecs ∇f=\vecs{F} is the radius of convergence; if the power series only
parametric equations of a line | the set of converges at x=a, the radius of convergence is R=0; if
equations x=x_0+ta, y=y_0+tb, and z=z_0+tc power function | a function of the form f(x)=x^n the power series converges for all real numbers x, the
describing the line with direction vector v=⟨a,b,c⟩ for any positive integer n≥1 radius of convergence is R=∞
passing through point (x_0,y_0,z_0)
power law for limits | the limit law \lim_{x→a} radius of curvature | the reciprocal of the
partial derivative | a derivative of a function of (f(x))^n=(\lim_{x→a}f(x))^n=L^n \nonumber for curvature
more than one independent variable in which all the every positive integer n
variables but one are held constant radius of gyration | the distance from an object’s
power reduction formula | a rule that allows an center of mass to its axis of rotation
partial differential equation | an equation that integral of a power of a trigonometric function to be
involves an unknown function of more than one exchanged for an integral involving a lower power range | the set of outputs for a function
independent variable and one or more of its partial
power rule | the derivative of a power function is a ratio test | for a series \displaystyle
derivatives
function in which the power on x becomes the \sum^∞_{n=1}a_n with nonzero terms, let
partial fraction decomposition | a technique coefficient of the term and the power on x in the \displaystyle ρ=\lim_{n→∞}|a_{n+1}/a_n|; if 0≤ρ<1,
used to break down a rational function into the sum of derivative decreases by 1: If n is an integer, then the series converges absolutely; if ρ>1, the series
simple rational functions \dfrac{d}{dx}\left(x^n\right)=nx^{n−1} diverges; if ρ=1, the test is inconclusive
partial sum | the kth partial sum of the infinite power series | a series of the form rational function | a function of the form
series \displaystyle \sum^∞_{n=1}a_n is the finite sum \sum_{n=0}^∞c_nx^n is a power series centered at f(x)=p(x)/q(x), where p(x) and q(x) are polynomials
\displaystyle x=0; a series of the form \sum_{n=0}^∞c_n(x−a)^n is recurrence relation | a recurrence relation is a
S_k=\sum_{n=1}^ka_n=a_1+a_2+a_3+⋯+a_k a power series centered at x=a relationship in which a term a_n in a sequence is
particular solution | member of a family of principal unit normal vector | a vector defined in terms of earlier terms in the sequence
solutions to a differential equation that satisfies a orthogonal to the unit tangent vector, given by the
particular initial condition
region | an open, connected, nonempty subset of
formula \frac{\vecs T′(t)}{‖\vecs T′(t)‖} \mathbb{R}^2
particular solution | a solution y_p(x) of a principal unit tangent vector | a unit vector
differential equation that contains no arbitrary
regular parameterization | parameterization
tangent to a curve C \vecs r(u,v) = \langle x(u,v), \, y(u,v), \, z(u,v)\rangle
constants
product law for limits | the limit law \lim_{x→a} such that r_u \times r_v is not zero for point (u,v) in
partition | a set of points that divides an interval into (f(x)⋅g(x))=\lim_{x→a}f(x)⋅\lim_{x→a}g(x)=L⋅M the parameter domain
subintervals \nonumber regular partition | a partition in which the
percentage error | the relative error expressed as a subintervals all have the same width
percentage

6 https://math.libretexts.org/@go/page/51385
related rates | are rates of change associated with scalar equation of a plane | the equation solution curve | a curve graphed in a direction field
two or more related quantities that are changing over a(x−x_0)+b(y−y_0)+c(z−z_0)=0 used to describe a that corresponds to the solution to the initial-value
time plane containing point P=(x_0,y_0,z_0) with normal problem passing through a given point in the direction
vector n=⟨a,b,c⟩ or its alternate form ax+by+cz+d=0, field
relative error | given an absolute error Δq for a where d=−ax_0−by_0−cz_0
particular quantity, \frac{Δq}{q} is the relative error. solution to a differential equation | a function
scalar line integral | the scalar line integral of a y=f(x) that satisfies a given differential equation
relative error | error as a percentage of the actual function f along a curve C with respect to arc length is
value, given by \text{relative error}=\left|\frac{A−B} the integral \displaystyle \int_C f\,ds, it is the integral space curve | the set of ordered triples (f(t),g(t),h(t))
{A}\right|⋅100\% \nonumber of a scalar function f along a curve in a plane or in together with their defining parametric equations
space; such an integral is defined in terms of a x=f(t), y=g(t) and z=h(t)
remainder estimate | for a series \displaystyle
\sum^∞_{n=}1a_n with positive terms a_n and a Riemann sum, as is a single-variable integral space-filling curve | a curve that completely
continuous, decreasing function f such that f(n)=a_n occupies a two-dimensional subset of the real plane
scalar multiplication | a vector operation that
for all positive integers n, the remainder \displaystyle
defines the product of a scalar and a vector speed | is the absolute value of velocity, that is, |v(t)|
R_N=\sum^∞_{n=1}a_n−\sum^N_{n=1}a_n satisfies
the following estimate: scalar projection | the magnitude of the vector is the speed of an object at time t whose velocity is
∫^∞_{N+1}f(x)\,dx<R_N<∫^∞_Nf(x)\,dx \nonumber projection of a vector given by v(t)

removable discontinuity | A removable secant | A secant line to a function f(x) at a is a line sphere | the set of all points equidistant from a given
point known as the center
discontinuity occurs at a point a if f(x) is discontinuous through the point (a,f(a)) and another point on the
at a, but \displaystyle \lim_{x→a}f(x) exists function; the slope of the secant line is given by spherical coordinate system | a way to describe
m_{sec}=\dfrac{f(x)−f(a)}{x−a} a location in space with an ordered triple (ρ,θ,φ),
reparameterization | an alternative
where ρ is the distance between P and the origin (ρ≠0),
parameterization of a given vector-valued function second derivative test | suppose f'(c)=0 and f'' is
θ is the same angle used to describe the location in
continuous over an interval containing c; if f''(c)>0,
restricted domain | a subset of the domain of a then f has a local minimum at c; if f''(c)<0, then f has a
cylindrical coordinates, and φ is the angle formed by
function f the positive z-axis and line segment \bar{OP}, where
local maximum at c; if f''(c)=0, then the test is
O is the origin and 0≤φ≤π
riemann sum | an estimate of the area under the inconclusive
curve of the form A≈\displaystyle separable differential equation | any equation squeeze theorem | states that if f(x)≤g(x)≤h(x) for
\sum_{i=1}^nf(x^∗_i)Δx that can be written in the form y'=f(x)g(y) all x≠a over an open interval containing a and
\lim_{x→a}f(x)=L=\lim_ {x→a}h(x) where L is a real
right-endpoint approximation | the right- separation of variables | a method used to solve a number, then \lim_{x→a}g(x)=L
endpoint approximation is an approximation of the
separable differential equation
area of the rectangles under a curve using the right standard equation of a sphere | (x−a)^2+
endpoint of each subinterval to construct the vertical sequence | an ordered list of numbers of the form (y−b)^2+(z−c)^2=r^2 describes a sphere with center
sides of each rectangle \displaystyle a_1,a_2,a_3,… is a sequence (a,b,c) and radius r
right-hand rule | a common way to define the sigma notation | (also, summation notation) the standard form | the form of a first-order linear
orientation of the three-dimensional coordinate system; Greek letter sigma (Σ) indicates addition of the values; differential equation obtained by writing the
when the right hand is curved around the z-axis in such the values of the index above and below the sigma differential equation in the form y'+p(x)y=q(x)
a way that the fingers curl from the positive x-axis to indicate where to begin the summation and where to
the positive y-axis, the thumb points in the direction of end it
standard form | an equation of a conic section
showing its properties, such as location of the vertex or
the positive z-axis
simple curve | a curve that does not cross itself lengths of major and minor axes
RLC series circuit | a complete electrical path
consisting of a resistor, an inductor, and a capacitor; a
simple harmonic motion | motion described by standard unit vectors | unit vectors along the
the equation x(t)=c_1 \cos (ωt)+c_2 \sin (ωt), as coordinate axes: \hat{\mathbf i}=⟨1,0⟩,\, \hat{\mathbf
second-order, constant-coefficient differential equation
exhibited by an undamped spring-mass system in j}=⟨0,1⟩
can be used to model the charge on the capacitor in an
which the mass continues to oscillate indefinitely
RLC series circuit standard-position vector | a vector with initial
rolle’s theorem | if f is continuous over [a,b] and simply connected region | a region that is point (0,0)
differentiable over (a,b), and if f(a)=f(b), then there connected and has the property that any closed curve
that lies entirely inside the region encompasses points
steady-state solution | a solution to a
exists c∈(a,b) such that f′(c)=0 nonhomogeneous differential equation related to the
that are entirely inside the region
forcing function; in the long term, the solution
root function | a function of the form f(x)=x^{1/n}
for any integer n≥2
Simpson’s rule | a rule that approximates approaches the steady-state solution
\displaystyle ∫^b_af(x)\,dx using the area under a
root law for limits | the limit law piecewise quadratic function. The approximation S_n
step size | the increment hh that is added to the xx
value at each step in Euler’s Method
\lim_{x→a}\sqrt[n]{f(x)}=\sqrt[n] to \displaystyle ∫^b_af(x)\,dx is given by
{\lim_{x→a}f(x)}=\sqrt[n]{L} for all L if n is odd and S_n=\frac{Δx} Stokes’ theorem | relates the flux integral over a
for L≥0 if n is even {3}\big(f(x_0)+4\,f(x_1)+2\,f(x_2)+4\,f(x_3)+2\,f(x_4 surface S to a line integral around the boundary C of
)+⋯+2\,f(x_{n−2})+4\,f(x_{n−1})+f(x_n)\big). the surface S
root test | for a series
\displaystyle
\nonumber
\sum^∞_{n=1}a_n, let \displaystyle stream function | if \vecs F=⟨P,Q⟩ is a source-free
ρ=\lim_{n→∞}\sqrt[n]{|a_n|}; if 0≤ρ<1, the series skew lines | two lines that are not parallel but do not vector field, then stream function g is a function such
converges absolutely; if ρ>1, the series diverges; if intersect that P=g_y and Q=−g_x
ρ=1, the test is inconclusive
slicing method | a method of calculating the sum law for limits | The limit law \lim_{x→a}
rose | graph of the polar equation r=a\cos 2θ or volume of a solid that involves cutting the solid into (f(x)+g(x))=\lim_{x→a}f(x)+\lim_{x→a}g(x)=L+M
r=a\sin 2θfor a positive constant a pieces, estimating the volume of each piece, then
adding these estimates to arrive at an estimate of the sum rule | the derivative of the sum of a function f
rotational field | a vector field in which the vector total volume; as the number of slices goes to infinity, and a function g is the same as the sum of the
at point (x,y) is tangent to a circle with radius derivative of f and the derivative of g: \dfrac{d}
this estimate becomes an integral that gives the exact
r=\sqrt{x^2+y^2}; in a rotational field, all vectors flow {dx}\big(f(x)+g(x)\big)=f′(x)+g′(x)
value of the volume
either clockwise or counterclockwise, and the
magnitude of a vector depends only on its distance slope | the change in y for each unit change in x surface | the graph of a function of two variables,
from the origin z=f(x,y)
slope-intercept form | equation of a linear
rulings | parallel lines that make up a cylindrical function indicating its slope and y-intercept surface area | the surface area of a solid is the total
surface area of the outer layer of the object; for objects such as
smooth | curves where the vector-valued function cubes or bricks, the surface area of the object is the
saddle point | given the function z=f(x,y), the point \vecs r(t) is differentiable with a non-zero derivative sum of the areas of all of its faces
(x_0,y_0,f(x_0,y_0)) is a saddle point if both
f_x(x_0,y_0)=0 and f_y(x_0,y_0)=0, but f does not
solid of revolution | a solid generated by revolving surface area | the area of surface S given by the
a region in a plane around a line in that plane surface integral \iint_S \,dS \nonumber
have a local extremum at (x_0,y_0)
scalar | a real number

7 https://math.libretexts.org/@go/page/51385
surface independent | flux integrals of curl vector term-by-term differentiation of a power trigonometric integral | an integral involving
fields are surface independent if their evaluation does series | a technique for evaluating the derivative of a powers and products of trigonometric functions
not depend on the surface but only on the boundary of power series \displaystyle \sum_{n=0}^∞c_n(x−a)^n
the surface by evaluating the derivative of each term separately to trigonometric substitution | an integration
create the new power series \displaystyle technique that converts an algebraic integral
surface integral | an integral of a function over a \sum_{n=1}^∞nc_n(x−a)^{n−1} containing expressions of the form \sqrt{a^2−x^2},
surface \sqrt{a^2+x^2}, or \sqrt{x^2−a^2} into a
term-by-term integration of a power series | trigonometric integral
surface integral of a scalar-valued function | a technique for integrating a power series \displaystyle
a surface integral in which the integrand is a scalar \sum_{n=0}^∞c_n(x−a)^n by integrating each term triple integral | the triple integral of a continuous
function separately to create the new power series \displaystyle function f(x,y,z) over a rectangular solid box B is the
C+\sum_{n=0}^∞c_n\dfrac{(x−a)^{n+1}}{n+1} limit of a Riemann sum for a function of three
surface integral of a vector field | a surface
variables, if this limit exists
integral in which the integrand is a vector field terminal point | the endpoint of a vector
triple integral in cylindrical coordinates | the
symmetric equations of a line | the equations theorem of Pappus for volume | this theorem limit of a triple Riemann sum, provided the following
\dfrac{x−x_0}{a}=\dfrac{y−y_0}{b}=\dfrac{z−z_0} states that the volume of a solid of revolution formed limit exists: lim_{l,m,n\rightarrow\infty}
{c} describing the line with direction vector v=⟨a,b,c⟩ by revolving a region around an external axis is equal \sum_{i=1}^l \sum_{j=1}^m \sum_{k=1}^n
passing through point (x_0,y_0,z_0) to the area of the region multiplied by the distance f(r_{ijk}^*, \theta_{ijk}^*, s_{ijk}^*) r_{ijk}^*
symmetry about the origin | the graph of a traveled by the centroid of the region \Delta r \Delta \theta \Delta z \nonumber
function f is symmetric about the origin if (−x,−y) is three-dimensional rectangular coordinate
on the graph of f whenever (x,y) is on the graph
triple integral in spherical coordinates | the
system | a coordinate system defined by three lines limit of a triple Riemann sum, provided the following
symmetry about the y-axis | the graph of a that intersect at right angles; every point in space is limit exists: lim_{l,m,n\rightarrow\infty}
function f is symmetric about the y-axis if (−x,y) is on described by an ordered triple (x,y,z) that plots its \sum_{i=1}^l \sum_{j=1}^m \sum_{k=1}^n
the graph of f whenever (x,y) is on the graph location relative to the defining axes f(\rho_{ijk}^*, \theta_{ijk}^*, \varphi_{ijk}^*)
threshold population | the minimum population (\rho_{ijk}^*)^2 \sin \, \varphi \Delta \rho \Delta \theta
symmetry principle | the symmetry principle \Delta \varphi \nonumber
states that if a region R is symmetric about a line I, that is necessary for a species to survive
then the centroid of R lies on I total area | total area between a function and the x- Type I | a region D in the xy- plane is Type I if it lies
axis is calculated by adding the area above the x-axis between two vertical lines and the graphs of two
table of values | a table containing a list of inputs continuous functions g_1(x) and g_2(x)
and their corresponding outputs and the area below the x-axis; the result is the same as
the definite integral of the absolute value of the Type II | a region D in the xy-plane is Type II if it
tangent | A tangent line to the graph of a function at function lies between two horizontal lines and the graphs of two
a point (a,f(a)) is the line that secant lines through continuous functions h_1(y) and h_2(h)
(a,f(a)) approach as they are taken through points on total differential | the total differential of the
the function with x-values that approach a; the slope of function f(x,y) at (x_0,y_0) is given by the formula unbounded sequence | a sequence that is not
the tangent line to a graph at a measures the rate of dz=f_x(x_0,y_0)dx+fy(x_0,y_0)dy bounded is called unbounded
change of the function at a trace | the intersection of a three-dimensional surface unit vector | a vector with magnitude 1
tangent line approximation (linearization) | with a coordinate plane
since the linear approximation of f at x=a is defined unit vector field | a vector field in which the
transcendental function | a function that cannot magnitude of every vector is 1
using the equation of the tangent line, the linear be expressed by a combination of basic arithmetic
approximation of f at x=a is also known as the tangent operations upper sum | a sum obtained by using the maximum
line approximation to f at x=a value of f(x) on each subinterval
transformation | a function that transforms a
tangent plane | given a function f(x,y) that is region GG in one plane into a region RR in another variable of integration | indicates which variable
differentiable at a point (x_0,y_0), the equation of the plane by a change of variables you are integrating with respect to; if it is x, then the
tangent plane to the surface z=f(x,y) is given by function in the integrand is followed by dx
z=f(x_0,y_0)+f_x(x_0,y_0)(x−x_0)+f_y(x_0,y_0) transformation of a function | a shift, scaling,
(y−y_0) or reflection of a function vector | a mathematical object that has both
magnitude and direction
tangent vector | to \vecs{r}(t) at t=t_0 any vector trapezoidal rule | a rule that approximates
\vecs v such that, when the tail of the vector is placed \displaystyle ∫^b_af(x)\,dx using the area of trapezoids. vector addition | a vector operation that defines the
at point \vecs r(t_0) on the graph, vector \vecs{v} is The approximation T_n to \displaystyle ∫^b_af(x)\,dx sum of two vectors
tangent to curve C is given by T_n=\frac{Δx}{2}\big(f(x_0)+2\, vector difference | the vector difference \vecs{v}−
f(x_1)+2\, f(x_2)+⋯+2\, f(x_{n−1})+f(x_n)\big). \vecs{w} is defined as \vecs{v}+(−
tangential component of acceleration | the \nonumber
coefficient of the unit tangent vector \vecs T when the \vecs{w})=\vecs{v}+(−1)\vecs{w}
acceleration vector is written as a linear combination tree diagram | illustrates and derives formulas for vector equation of a line | the equation \vecs
of \vecs T and \vecs N the generalized chain rule, in which each independent
r=\vecs r_0+t\vecs v used to describe a line with
variable is accounted for
Taylor polynomials | the n^{\text{th}}-degree direction vector \vecs v=⟨a,b,c⟩ passing through point
Taylor polynomial for f at x=a is p_n(x)=f(a)+f′(a) triangle inequality | If a and b are any real P=(x_0,y_0,z_0), where \vecs r_0=⟨x_0,y_0,z_0⟩, is
(x−a)+\dfrac{f''(a)}{2!}(x−a)^2+⋯+\dfrac{f^{(n)} numbers, then |a+b|≤|a|+|b| the position vector of point P
(a)}{n!}(x−a)^n vector equation of a plane | the equation \vecs
triangle inequality | the length of any side of a
Taylor series | a power series at a that converges to triangle is less than the sum of the lengths of the other n⋅\vecd{PQ}=0, where P is a given point in the plane,
a function f on some open interval containing a. two sides Q is any point in the plane, and \vecs n is a normal
vector of the plane
Taylor’s theorem with remainder | for a triangle method | a method for finding the sum of
function f and the n^{\text{th}}-degree Taylor two vectors; position the vectors so the terminal point vector field | measured in ℝ^2, an assignment of a
polynomial for f at x=a, the remainder R_n(x)=f(x) of one vector is the initial point of the other; these vector \vecs{F}(x,y) to each point (x,y) of a subset D
−p_n(x) satisfies R_n(x)=\dfrac{f^{(n+1)}(c)} vectors then form two sides of a triangle; the sum of of ℝ^2; in ℝ^3, an assignment of a vector \vecs{F}
{(n+1)!}(x−a)^{n+1} for somec between x and a; if the vectors is the vector that forms the third side; the (x,y,z) to each point (x,y,z) of a subset D of ℝ^3
there exists an interval I containing a and a real initial point of the sum is the initial point of the first
vector line integral | the vector line integral of
number M such that ∣f^{(n+1)}(x)∣≤M for all x in I, vector; the terminal point of the sum is the terminal
vector field \vecs F along curve C is the integral of the
then |R_n(x)|≤\dfrac{M}{(n+1)!}|x−a|^{n+1} point of the second vector
dot product of \vecs F with unit tangent vector \vecs T
telescoping series | a telescoping series is one in trigonometric functions | functions of an angle of C with respect to arc length, ∫_C \vecs F·\vecs T\,
which most of the terms cancel in each of the partial defined as ratios of the lengths of the sides of a right ds; such an integral is defined in terms of a Riemann
sums triangle sum, similar to a single-variable integral

term | the number \displaystyle a_n in the sequence trigonometric identity | an equation involving vector parameterization | any representation of a
\displaystyle {a_n} is called the \displaystyle nth term trigonometric functions that is true for all angles θ for plane or space curve using a vector-valued function
of the sequence which the functions in the equation are defined
vector projection | the component of a vector that
follows a given direction

8 https://math.libretexts.org/@go/page/51385
vector sum | the sum of two vectors, \vecs{v} and vertex | a vertex is an extreme point on a conic work | the amount of energy it takes to move an
\vecs{w}, can be constructed graphically by placing section; a parabola has one vertex at its turning point. object; in physics, when a force is constant, work is
the initial point of \vecs{w} at the terminal point of An ellipse has two vertices, one at each end of the expressed as the product of force and distance
\vecs{v}; then the vector sum \vecs{v}+\vecs{w} is major axis; a hyperbola has two vertices, one at the
the vector with an initial point that coincides with the turning point of each branch
work done by a force | work is generally thought
of as the amount of energy it takes to move an object;
initial point of \vecs{v}, and with a terminal point that
coincides with the terminal point of \vecs{w}
vertical asymptote | A function has a vertical if we represent an applied force by a vector \vecs{ F}
asymptote at x=a if the limit as x approaches a from and the displacement of an object by a vector \vecs{
vector-valued function | a function of the form the right or left is infinite s}, then the work done by the force is the dot product
\vecs r(t)=f(t)\hat{\mathbf{i}}+g(t)\hat{\mathbf{j}} or of \vecs{ F} and \vecs{ s}.
\vecs vertical line test | given the graph of a function,
r(t)=f(t)\hat{\mathbf{i}}+g(t)\hat{\mathbf{j}}+h(t)\ha every vertical line intersects the graph, at most, once zero vector | the vector with both initial point and
t{\mathbf{k}},where the component functions f, g, terminal point (0,0)
vertical trace | the set of ordered triples (c,y,z) that
and h are real-valued functions of the parameter t. solves the equation f(c,y)=z for a given constant x=c or zeros of a function | when a real number x is a
the set of ordered triples (x,d,z) that solves the zero of a function f,\;f(x)=0
velocity vector | the derivative of the position
equation f(x,d)=z for a given constant y=d
vector δ ball | all points in \mathbb{R}^3 lying at a distance
washer method | a special case of the slicing of less than δ from (x_0,y_0,z_0)
method used with solids of revolution when the slices
are washers
δ disk | an open disk of radius δ centered at point
(a,b)

9 https://math.libretexts.org/@go/page/51385
Detailed Licensing
Overview
Title: Calculus 3e (Apex)
Webpages: 127
Applicable Restrictions: Noncommercial
All licenses found:
CC BY-NC 3.0: 89% (113 pages)
Undeclared: 11% (14 pages)

By Page
Calculus 3e (Apex) - CC BY-NC 3.0 4: Applications of the Derivative - CC BY-NC 3.0
Front Matter - Undeclared 4.1: Newton's Method - CC BY-NC 3.0
TitlePage - Undeclared 4.2: Related Rates - CC BY-NC 3.0
InfoPage - Undeclared 4.3: Optimization - CC BY-NC 3.0
Table of Contents - Undeclared 4.4: Differentials - CC BY-NC 3.0
Licensing - Undeclared 4.E: Applications of Derivatives (Exercises) - CC BY-
NC 3.0
1: Limits - CC BY-NC 3.0
5: Integration - CC BY-NC 3.0
1.1: An Introduction to Limits - CC BY-NC 3.0
1.2: Epsilon-Delta Definition of a Limit - CC BY-NC 5.1: Antiderivatives and Indefinite Integration - CC
3.0 BY-NC 3.0
1.3: Finding Limits Analytically - CC BY-NC 3.0 5.2: The Definite Integral - CC BY-NC 3.0
1.4: One Sided Limits - CC BY-NC 3.0 5.3: Riemann Sums - CC BY-NC 3.0
1.5: Continuity - CC BY-NC 3.0 5.4: The Fundamental Theorem of Calculus - CC BY-
1.6: Limits Involving Infinity - CC BY-NC 3.0 NC 3.0
1.E: Applications of Limits (Exercises) - CC BY-NC 5.5: Numerical Integration - CC BY-NC 3.0
3.0 5.E: Applications of Integration (Exercises) - CC BY-
2: Derivatives - CC BY-NC 3.0 NC 3.0

2.1: Instantaneous Rates of Change- The Derivative - 6: Techniques of Integration - CC BY-NC 3.0
CC BY-NC 3.0 Front Matter - Undeclared
2.2: Interpretations of the Derivative - CC BY-NC 3.0 TitlePage - Undeclared
2.3: Basic Differentiation Rules - CC BY-NC 3.0 InfoPage - Undeclared
2.4: The Product and Quotient Rules - CC BY-NC 3.0 6.1: Substitution - CC BY-NC 3.0
2.5: The Chain Rule - CC BY-NC 3.0 6.2: Integration by Parts - CC BY-NC 3.0
2.6: Implicit Differentiation - CC BY-NC 3.0 6.3: Trigonometric Integrals - CC BY-NC 3.0
2.7: Derivatives of Inverse Functions - CC BY-NC 3.0 6.4: Trigonometric Substitution - CC BY-NC 3.0
2.E: Applications of Derivatives(Exercises) - CC BY- 6.5: Partial Fraction Decomposition - CC BY-NC 3.0
NC 3.0 6.6: Hyperbolic Functions - CC BY-NC 3.0
3: The Graphical Behavior of Functions - CC BY-NC 3.0 6.7: L'Hopital's Rule - CC BY-NC 3.0
3.1: Extreme Values - CC BY-NC 3.0 6.8: Improper Integration - CC BY-NC 3.0
3.2: The Mean Value Theorem - CC BY-NC 3.0 6.E: Applications of Antidifferentiation (Exercises) -
3.3: Increasing and Decreasing Functions - CC BY- CC BY-NC 3.0
NC 3.0 Back Matter - Undeclared
3.4: Concavity and the Second Derivative - CC BY- Index - Undeclared
NC 3.0 7: Applications of Integration - CC BY-NC 3.0
3.5: Curve Sketching - CC BY-NC 3.0
7.1: Area Between Curves - CC BY-NC 3.0
3.E: Applications of the Graphical Behavior of
Functions(Exercises) - CC BY-NC 3.0

1 https://math.libretexts.org/@go/page/115412
7.2: Volume by Cross-Sectional Area- Disk and 11.4: Unit Tangent and Normal Vectors - CC BY-NC
Washer Methods - CC BY-NC 3.0 3.0
7.3: The Shell Method - CC BY-NC 3.0 11.5: The Arc Length Parameter and Curvature - CC
7.4: Arc Length and Surface Area - CC BY-NC 3.0 BY-NC 3.0
7.5: Work - CC BY-NC 3.0 11.E: Applications of Vector Valued Functions
7.6: Fluid Forces - CC BY-NC 3.0 (Exercises) - CC BY-NC 3.0
7.E: Applications of Integration (Exercises) - CC BY- 12: Functions of Several Variables - CC BY-NC 3.0
NC 3.0 12.1: Introduction to Multivariable Functions - CC
8: Sequences and Series - CC BY-NC 3.0 BY-NC 3.0
8.1: Sequences - CC BY-NC 3.0 12.2: Limits and Continuity of Multivariable
8.2: Infinite Series - CC BY-NC 3.0 Functions - CC BY-NC 3.0
8.3: Integral and Comparison Tests - CC BY-NC 3.0 12.3: Partial Derivatives - CC BY-NC 3.0
8.4: Ratio and Root Tests - CC BY-NC 3.0 12.4: Differentiability and the Total Differential - CC
8.5: Alternating Series and Absolute Convergence - BY-NC 3.0
CC BY-NC 3.0 12.5: The Multivariable Chain Rule - CC BY-NC 3.0
8.6: Power Series - CC BY-NC 3.0 12.6: Directional Derivatives - CC BY-NC 3.0
8.7: Taylor Polynomials - CC BY-NC 3.0 12.7: Tangent Lines, Normal Lines, and Tangent
8.8: Taylor Series - CC BY-NC 3.0 Planes - CC BY-NC 3.0
8.E: Applications of Sequences and Series 12.8: Extreme Values - CC BY-NC 3.0
(Exercises) - CC BY-NC 3.0 12.E: Applications of Functions of Several Variables
9: Curves in the Plane - CC BY-NC 3.0 (Exercises) - CC BY-NC 3.0
9.1: Conic Sections - CC BY-NC 3.0 13: Multiple Integration - CC BY-NC 3.0
9.2: Parametric Equations - CC BY-NC 3.0 13.1: Iterated Integrals and Area - CC BY-NC 3.0
9.3: Calculus and Parametric Equations - CC BY-NC 13.2: Double Integration and Volume - CC BY-NC 3.0
3.0 13.3: Double Integration with Polar Coordinates - CC
9.4: Introduction to Polar Coordinates - CC BY-NC BY-NC 3.0
3.0 13.4: Center of Mass - CC BY-NC 3.0
9.5: Calculus and Polar Functions - CC BY-NC 3.0 13.5: Surface Area - CC BY-NC 3.0
9.E: Applications of Curves in a Plane (Exercises) - 13.6: Volume Between Surfaces and Triple
CC BY-NC 3.0 Integration - CC BY-NC 3.0
10: Vectors - CC BY-NC 3.0 13.E: Applications of Multiple Integration (Exercises)
10.1: Introduction to Cartesian Coordinates in Space - - CC BY-NC 3.0
CC BY-NC 3.0 14: Appendix - CC BY-NC 3.0
10.2: An Introduction to Vectors - CC BY-NC 3.0 14.1: Section 1- - CC BY-NC 3.0
10.3: The Dot Product - CC BY-NC 3.0 14.2: Section 2- - CC BY-NC 3.0
10.4: The Cross Product - CC BY-NC 3.0 14.3: Section 3- - CC BY-NC 3.0
10.5: Lines - CC BY-NC 3.0 14.4: Section 4- - CC BY-NC 3.0
10.6: Planes - CC BY-NC 3.0 14.5: Section 5- - CC BY-NC 3.0
10.E: Applications of Vectors (Exercises) - CC BY- 14.6: Section 6- - CC BY-NC 3.0
NC 3.0 Back Matter - Undeclared
11: Vector-Valued Functions - CC BY-NC 3.0 Index - Undeclared
11.1: Vector–Valued Functions - CC BY-NC 3.0 Glossary - Undeclared
11.2: Calculus and Vector-Valued Functions - CC BY- Detailed Licensing - Undeclared
NC 3.0
11.3: The Calculus of Motion - CC BY-NC 3.0

2 https://math.libretexts.org/@go/page/115412

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy