You are on page 1of 261

Unit 1

Graphing Functions, Domain and Range


Structure
1.1 Graphing Functions

1.2 Graphs of the Toolkit Functions

1.2.1 Domain and Range from Graphs

1.2.2 Domains and Ranges of the Toolkit Functions

1.3 Piecewise Functions

1.4 Compositions of Functions

1.5 Composition using Formulas

Self-Assessment Questions

References

Licenses and Attributions

The text is adapted by Symbiosis Centre for Distance Learning under a Creative Commons Attribution-ShareAlike 4.0
International (CC BY-SA 4.0) as requested by the work’s creator or licensees. This license is available at
https://creativecommons.org/licenses/by-sa/4.0/
Objectives
After going through this unit, you will be able to:

 Describe the graphing function with the help of graphs


 Recognize the importance of graphing function

1.1 GRAPHING FUNCTIONS


In this text, we will be exploring functions—the shapes of their graphs, their unique features, their equations, and
how to solve problems with them. When learning to read, we start with the alphabet. When learning to do arithmetic,
we start with numbers. When working with functions, it is similarly helpful to have a base set of elements to build
from. We call these our “toolkit of functions”—a set of basic named functions for which we know the graph, equation,
and special features.

For these definitions we will use x as the input variable and f(x) as the output variable.

https://youtu.be/sW9-zBeQpCU

You will see these toolkit functions, combinations of toolkit functions, their graphs and their transformations
frequently throughout this book. In order to successfully follow along later in the book, it will be very helpful if you
can recognize these toolkit functions and their features quickly by name, equation, graph and basic table values.

Not every important equation can be written as y = f(x). An example of this is the equation of a circle. Recall the
distance formula for the distance between two points:

A circle with radius r with center at (h, k) can be described as all points (x, y) a distance of r from the center, so using

the distance formula, , giving the equation of a circle.

Equation of a Circle

A circle with radius r with center (h, k) has equation


1.2 GRAPHS OF THE TOOLKIT FUNCTIONS

Important Topics of this Section

 Toolkit Functions

Domain and Range

One of our main goals in mathematics is to model the real world with mathematical functions. In doing so, it is
important to keep in mind the limitations of those models we create.

This table shows a relationship between circumference and height of a tree as it grows.

Circumference, c 1.7 2.5 5.5 8.2 13.7


Height, h 24.5 31 45.2 54.6 92.1

While there is a strong relationship between the two, it would certainly be ridiculous to talk about a tree with a
circumference of –3 feet, or a height of 3000 feet. When we identify limitations on the inputs and outputs of a
function, we are determining the domain and range of the function.
Domain and Range
 Domain: The set of possible input values to a function
 Range: The set of possible output values of a function

Example 1.1
Using the tree table above, determine a reasonable domain and range.

We could combine the data provided with our own experiences and reason to approximate
the domain and range of the function h = f(c). For the domain, possible values for the input
circumference c, it doesn’t make sense to have negative values, so c > 0. We could make an
educated guess at a maximum reasonable value, or look up that the maximum circumference
measured is about 119 feet.1 With this information we would say a reasonable domain is
0 < c ≤ 119 feet.

Similarly for the range, it doesn’t make sense to have negative heights, and the maximum
height of a tree could be looked up to be 379 feet, so a reasonable range is 0 < h ≤ 379 feet.

Example 1.2
When sending a letter through the United States Postal Service, the price depends upon the
weight of the letter, as shown in the table below. Determine the domain and range.

Letters
Weight not Over Price
1 ounce $0.44
2 ounces $0.61
3 ounces $0.78
3.5 ounces $0.95

Suppose we notate Weight by w and Price by p, and set up a function named P, where
Price, p is a function of Weight, w. p = P(w)

Since acceptable weights are 3.5 ounces or less, and negative weights don’t make sense, the
domain would be. Technically 0 could be included in the domain, but logically it would mean
we are mailing nothing, so it doesn’t hurt to leave it out.

Since possible prices are from a limited set of values, we can only define the range of this
function by listing the possible values. The range is p = $0.44, $0.61, $0.78, or $0.95.

Activity 1
The population of a small town in the year 1960 was 100 people. Since then the population has grown
to 1400 people reported during the 2010 census. Choose descriptive variables for your input and
output and use interval notation to write the domain and range.

1
http://en.wikipedia.org/wiki/Tree, retrieved July 19, 2010
Notation
In the previous examples, we used inequalities to describe the domain and range of the functions. This is one way to
describe intervals of input and output values, but is not the only way. Let us take a moment to discuss notation for
domain and range.

Using inequalities, such as 0 < c ≤ 163, 0 < w ≤ 3.5, and 0 < h ≤ 379 imply that we are interested in all values between
the low and high values, including the high values in these examples.

However, occasionally we are interested in a specific list of numbers like the range for the price to send
letters, p = $0.44, $0.61, $0.78, or $0.95. These numbers represent a set of specific values: {0.44, 0.61, 0.78, 0.95}

Representing values as a set, or giving instructions on how a set is built, leads us to another type of notation to
describe the domain and range.

Suppose we want to describe the values for a variable x that are 10 or greater, but less than 30. In inequalities, we
would write 10 ≤ x < 30.

When describing domains and ranges, we sometimes extend this into set-builder notation, which would look like
this: {x|10 ≤ x < 30}. The curly brackets {} are read as “the set of,” and the vertical bar, | is read as “such that,” so
altogether we would read {x|10 ≤ x < 30} as “the set of x-values such that 10 is less than or equal to x and x is less
than 30.”

When describing ranges in set-builder notation, we could similarly write something like {f(x)|0 <f(x) < 100}, or if the
output had its own variable, we could use it. So for our tree height example above, we could write for the range
{h|0 < h ≤ 379}. In set-builder notation, if a domain or range is not limited, we could write {t|t is a real number}, or
{t|t ∈°}, read as “the set of t-values such that t is an element of the set of real numbers.

A more compact alternative to set-builder notation is interval notation, in which intervals of values are referred to
by the starting and ending values. Curved parentheses are used for “strictly less than,” and square brackets are used
for “less than or equal to.” Since infinity is not a number, we can’t include it in the interval, so we always use curved
parentheses with ∞ and –∞. The table below will help you see how inequalities correspond to set-builder notation
and interval notation:

Inequality Set Builder Notation Interval notation


5 < h ≤ 10 {h|5 < h ≤ 10} (5,10]
5 ≤ h < 10 {h|5 ≤ h < 10} [5,10)
5 < h < 10 {h|5 < h < 10} (5,10)
h < 10 {h|h < 10} (–∞,10)
h ≥ 10 {h|h ≥ 10} [10,∞)
all real numbers {h|h ∈°} (–∞,∞)

To combine two intervals together, using inequalities or set-builder notation we can use the word “or.” In interval
notation, we use the union symbol, ∪, to combine two unconnected intervals together.

Example 1.3
Describe the intervals of values shown on the line graph below using set builder and interval
notations.
To describe the values, x, that lie in the intervals shown above we would say, “x is a real
number greater than or equal to 1 and less than or equal to 3, or a real number greater than
5.”

 As an inequality it is: 1 ≤ x ≤ 3 or x > 5


 In set builder notation: {x|1 ≤ x ≤ 3 or x > 5}
 In interval notation: [1,3]∪(5,∞)

Remember when writing or reading interval notation:

 Using a square bracket [ means the start value is included in the set
 Using a parenthesis ( means the start value is not included in the set

1.2.1 Domain and Range from Graphs


We can also talk about domain and range based on graphs. Since domain refers to the set of possible input values,
the domain of a graph consists of all the input values shown on the graph. Remember that input values are almost
always shown along the horizontal axis of the graph. Likewise, since range is the set of possible output values, the
range of a graph we can see from the possible values along the vertical axis of the graph.

Be careful—if the graph continues beyond the window on which we can see the graph, the domain and range might
be larger than the values we can see.

Example 1.4
Determine the domain and range of the graph below.

In the graph above, the input quantity along the horizontal axis appears to be “year,” which
we could notate with the variable y. The output is “thousands of barrels of oil per day,” which
we might notate with the variable b, for barrels. The graph would likely continue to the left
and right beyond what is shown, but based on the portion of the graph that is shown to us,
we can determine the domain is 1975 ≤ y ≤ 2008, and the range is approximately
180 ≤ b ≤ 2010.

In interval notation, the domain would be [1975, 2008] and the range would be about [180,
2010]. For the range, we have to approximate the smallest and largest outputs since they
don’t fall exactly on the grid lines.

Remember that, as in the previous example, x and y are not always the input and output
variables. Using descriptive variables is an important tool to remembering the context of the
problem.

Activity 2
Considering the graph given below, write the domain and range in interval notation

1.2.2 Domains and Ranges of the Toolkit Functions


We will now return to our set of toolkit functions to note the domain and range of each.

(1) Constant Function

f(x) = c

The domain here is not restricted; x can be anything. When this is the case we say the domain is all real numbers.
The outputs are limited to the constant value of the function.

Domain: (–∞,∞)

Range: [c]

Since there is only one output value, we list it by itself in square brackets.
(2) Identity Function

f(x) = x

Domain: (–∞,∞)

Range: (–∞,∞)

(3) Quadratic Function

f(x) = x2

Domain: (–∞,∞)

Range: [0,∞)

Multiplying a negative or positive number by itself can only yield a positive output.

(4) Cubic Function

f(x) = x3

Domain: (–∞,0)∪(0,∞)

Range: (–∞,0)∪(0,∞)

(5) Reciprocal

Domain: (–∞,0)∪(0,∞)

Range: (–∞,0)∪(0,∞)

We cannot divide by 0 so we must exclude 0 from the domain.

One divide by any value can never be 0, so the range will not include 0.

(6) Reciprocal Squared

Domain: (–∞,0)∪(0,∞)

Range: (0,∞)

We cannot divide by 0 so we must exclude 0 from the domain.

(7) Cube Root

Domain: (–∞,∞)

Range: (–∞,∞)
(8) Square Root

Domain: [0,∞)

Range: [0,∞)

When dealing with the set of real numbers we cannot take the square root of a negative number so the domain is
limited to 0 or greater.

(9) Absolute Value Function

f(x) = |x|

Domain: (–∞,∞)

Range: [0,∞)

Since absolute value is defined as a distance from 0, the output can only be greater than or equal to 0.

Example 1.5

1.3 PIECEWISE FUNCTIONS


In the toolkit functions, we introduced the absolute value function f(x) = |x|.

With a domain of all real numbers and a range of values greater than or equal to 0, the absolute value can be defined
as the magnitude or modulus of a number, a real number value regardless of sign, the size of the number, or the
distance from 0 on the number line. All of these definitions require the output to be greater than or equal to 0.

If we input 0, or a positive value the output is unchanged f(x) = x if x ≥ 0

If we input a negative value the sign must change from negative to positive.

f(x) = –x if x < 0 since multiplying a negative value by –1 makes it positive.

Since this requires two different processes or pieces, the absolute value function is often called the most basic
piecewise defined function.

A piecewise function is a function in which the formula used depends upon the domain the input lies in. We notate
this idea like:
Example 1.6
A museum charges $5 per person for a guided tour with a group of 1 to 9 people, or a fixed
$50 fee for 10 or more people in the group. Set up a function relating the number of people,
n, to the cost, C.

To set up this function, two different formulas would be needed. C = 5n would work
for n values under 10, and C = 50 would work for values of n ten or greater. Notating this:

Example 1.7
A cell phone company uses the function below to determine the cost,
C, in dollars for g gigabytes of data transfer.

 25 if 0 g 2
C(g)  
25  10( g  2) if g2

Find the cost of using 1.5 gigabytes of data, and the cost of using 4 gigabytes of data.

To find the cost of using 1.5 gigabytes of data,


C(1.5), we first look to see which piece of domain our input falls in. Since 1.5 is less than 2,
we use the first formula, giving C(1.5) = $25.

The find the cost of using 4 gigabytes of data,


C(4), we see that our input of 4 is greater than 2, so we’ll use the second formula.

C(4)=25+10(4−2)=$45

Example 1.8
Sketch a graph of the function

Since each of the component functions are from our library of Toolkit functions, we know
their shapes. We can imagine graphing each function, then limiting the graph to the indicated
domain. At the endpoints of the domain, we put open circles to indicate where the endpoint
is not included, due to a strictly-less-than inequality, and a closed circle where the endpoint
is included, due to a less-than-or-equal-to inequality.
For more information, please click on the below link to watch video

https://courses.lumenlearning.com/sanjacinto-finitemath1/chapter/find-a-quadratic-function-given-the-intercepts-
of-the-graph/

1.4 COMPOSITIONS OF FUNCTIONS


Suppose we wanted to calculate how much it costs to heat a house on a particular day of the year. The cost to heat
a house will depend on the average daily temperature, and the average daily temperature depends on the particular
day of the year. Notice how we have just defined two relationships: The temperature depends on the day, and the
cost depends on the temperature. Using descriptive variables, we can notate these two functions.

The first function, C(T), gives the cost C of heating a house when the average daily temperature is T degrees Celsius,
and the second, T(d), gives the average daily temperature of a particular city on day d of the year. If we wanted to
determine the cost of heating the house on the 5th day of the year, we could do this by linking our two functions
together, an idea called composition of functions. Using the function T(d), we could evaluate T(5) to determine the
average daily temperature on the 5th day of the year. We could then use that temperature as the input to the C(T)
function to find the cost to heat the house on the 5th day of the year: C(T(5)).

Definition: - “When the output of one function is used as the input of another, we call the entire operation
a composition of functions. We write f(g(x)), and read this as “f of g of x” or “f composed with g at x.”

An alternate notation for composition uses the composition operator \o\o is read “f of g of x” or “f composed
with g at x,” just like f(g(x)).

(f°g)=f(g(x))

Example 1.9
Suppose c(s) gives the number of calories burned doing s sit-ups, and s(t) gives the number
of sit-ups a person can do in t minutes. Interpret c(s(3)).

When we are asked to interpret, we are being asked to explain the meaning of the expression
in words. The inside expression in the composition is s(3). Since the input to the s function is
time, the 3 is representing 3 minutes, and s(3) is the number of sit-ups that can be done in 3
minutes. Taking this output and using it as the input to the c(s) function will gives us the
calories that can be burned by the number of sit-ups that can be done in 3 minutes.

Note that it is not important that the same variable be used for the output of the inside
function and the input to the outside function. However, it is essential that the units on the
output of the inside function match the units on the input to the outside function, if the units
are specified.

Example 1.10
Suppose f(x) gives miles that can be driven in x hours, and g(y) gives the gallons of gas used
in driving y miles. Which of these expressions is meaningful: f(g(y)) or g(f(x))?

The expression g(y) takes miles as the input and outputs a number of gallons. The
function f(x) is expecting a number of hours as the input; trying to give it a number of gallons
as input does not make sense. Remember the units have to match, and number of gallons
does not match number of hours, so the expression f(g(y)) is meaningless.
The expression f(x) takes hours as input and outputs a number of miles driven. The
function g(y) is expecting a number of miles as the input, so giving the output of
the f(x) function (miles driven) as an input value for g(y), where gallons of gas depends on
miles driven, does make sense. The expression g(f(x)) makes sense, and will give the number
of gallons of gas used, g, driving a certain number of miles, f(x), in x hours.

Activity 3

In a department store you see a sign that says 50% off of clearance merchandise, so final cost C depends on the
clearance price, p, according to the function C(p). Clearance price, p, depends on the original discount, d, given to
the clearance item, p(d).
Interpret C(p(d)).

Composition of Functions using Tables and Graphs

When working with functions given as tables and graphs, we can look up values for the functions using a provided
table or graph, as discussed in Section 1.1. We start evaluation from the provided input, and first evaluate the inside
function. We can then use the output of the inside function as the input to the outside function. To remember this,
always work from the inside out.

Example 1.11
Using the tables below, evaluate f(g(3)) and g(f(4))

x f(x) x g(x)
1 6 1 3
2 8 2 5
3 3 3 2
4 1 4 7

To evaluate f(g(3)), we start from the inside with the value 3. We then evaluate the inside
expression g(3) using the table that defines the function g: g(3) = 2. We can then use that
result as the input to the f function, so g(3) is replaced by the equivalent value 2 and we
get f(2). Then using the table that defines the function f, we find that f(2) = 8.

f(g(3)) = f(2) = 8

To evaluate g(f(4)), we first evaluate the inside expression f(4) using the first table: f(4) =
1. Then using the table for g we can evaluate: g(f(4)) = g(1) = 3

Example 1.12
Using the graphs below, evaluate f(g(1)).
To evaluate f(g(1)), we again start with the inside evaluation. We evaluate g(1) using the
graph of the g(x) function, finding the input of 1 on the horizontal axis and finding the output
value of the graph at that input. Here, g(1) = 3. Using this value as the input to
the f function, f(g(1) = f(3). We can then evaluate this by looking to the graph of
the f(x) function, finding the input of 3 on the horizontal axis, and reading the output value
of the graph at this input. Here, f(3) = 6, so f(g(1)) = 6.

1.5 COMPOSITION USING FORMULAS

When evaluating a composition of functions where we have either created or been given formulas, the concept of
working from the inside out remains the same. First we evaluate the inside function using the input value provided,
then use the resulting output as the input to the outside function.

Example 1.13
Given f(t) = t2 – t and h(x) = 3x + 2, evaluate f(h(1)).

Since the inside evaluation is h(1) we start by evaluating the h(x) function at
1: h(1) = 3(1) + 2 = 5

Then f(h(1)) = f(5), so we evaluate the f(t) function at an input of 5: f(h(1) = f(5) = 52 – 5 = 20

While we can compose the functions as above for each individual input value, sometimes it
would be really helpful to find a single formula which will calculate the result of a
composition f(g(x)). To do this, we will extend our idea of function evaluation. Recall that
when we evaluate a function like f(t) = t2 – t, we put whatever value is inside the parentheses
after the function name into the formula wherever we see the input variable.

Example 1.14
Using the same f(t) function from above, evaluate f(x + 2).

Everywhere in the formula for f where there was a t, we would replace it with the input
(x + 2). Since in the original formula the input t was squared in the first term, the entire input
x + 2 needs to be squared when we substitute, so we need to use grouping parentheses. To
avoid problems, it is advisable to always use parentheses around inputs.

f(x + 2) = (x + 2)2 – (x + 2)

We could simplify this expression further to f(x + 2) = x2 + 3x + 2 if we wanted to:


f(x+2)=(x+2)(x+2)−(x+2) Use the “FOIL” technique (first, outside, inside, last)

f(x+2)=x2+2x+2x+4−(x+2) Distribute the negative sign

f(x+2)=x2+2x+2x+4−x−2 Combine like terms

f(x+2)=x2+3x+2

Self-Assessment Questions
1) A city manager determines that the tax revenue,
R, in millions of dollars collected on a population of p thousand people is given by the
formula , where t is measured in years after 2010. Find a formula for the tax revenue as a
function of the year.

Since we want tax revenue as a function of the year, we want year to be our initial input, and revenue to be
our final output. To find revenue, we will first have to predict the city population, and then use that result as
the input to the tax function. So we need to find R(p(t)).

References

1. http://en.wikipedia.org/wiki/Tree, retrieved July 19, 2010.

2. http://www.usps.com/prices/first-class-mail-prices.htm, retrieved July 19, 2010.

3. https://www.pierce.ctc.edu/dist/tuition/ref/files/0910_tuition_rate.pdf, retrieved August 6, 2010.

Licenses and Attributions


CC Licensed Content, Shared Previously
 Chapter 1: Functions. Authored by: David Lippman and Melonie Rasmussen. Provided by: Open Text
Bookstore. Located at: http://www.opentextbookstore.com/precalc/1.4/Chapter%201.doc. Project:
Precalculus: An Investigation of Functions. License: CC BY-SA: Attribution-ShareAlike
 Introduction to Functions - Part 2. Authored by: James Sousa. Located at: https://youtu.be/sW9-
zBeQpCU. License: CC BY: Attribution

About MIT OpenCourseWare


MIT OpenCourseWare makes the materials used in the teaching of almost all of MIT's subjects available
on the Web, free of charge. With more than 2,400 courses available, OCW is delivering on the promise of
open sharing of knowledge. Learn more »
Massachusetts Institute of Technology logo and name. MIT Office of Digital Learning logo and name.
Open Education Consortium logo. Creative Commons logo with terms BY-NC-SA.
© 2001–2019
Massachusetts Institute of Technology
2 ATTRIBUTION
Licensing Terms of Use Accessibility Statement Contact
Supported by William & Flora Hewlett Foundation, Bill & Melinda Gates Foundation, Michelson 20MM
Foundation, Maxfield Foundation, Open Society Foundations, and Rice University. Powered by OpenStax
CNX.
Advanced Placement® and AP® are trademarks registered and/or owned by the College Board, which
was not involved in the production of, and does not endorse, this site.
Creative Commons © 1999-2019, Rice University. Except where otherwise noted, content created on this
site is licensed under a Creative Commons Attribution 4.0 License.
https://courses.lumenlearning.com/sanjacinto-finitemath1/chapter/reading-functions-and-function-ation-part-ii/
Unit 2
Polynomial and Rational Functions

Structure
2.1 Power Functions and Polynomial Functions
2.2 Quadratic Functions
2.3 Graphs of Polynomial Functions
2.4 Rational Functions
2.5 Inverse and Radical Functions
Self-Assessment Questions
Answers to Check your Progress
Attributes

The text is adapted by Symbiosis Centre for Distance Learning under a Creative Commons Attribution-
ShareAlike 4.0 International (CC BY-SA 4.0) as requested by the work’s creator or licensees. This license
is available at https://creativecommons.org/licenses/by-sa/4.0/
Objectives
After going through this unit, you will be able to:

 Describe Polynomial and Rational Functions


 Recognize the importance of graphs in Polynomial and Rational Functions

2.1 POWER FUNCTIONS AND POLYNOMIAL FUNCTIONS


A square is cut out of cardboard, with each side having length L. If we wanted to write a function for the
area of the square, with L as the input and the area as output, you may recall that the area of a rectangle
can be found by multiplying the length times the width. Since our shape is a square, the length & the
width are the same, giving the formula:

A( L)  L  L  L2

Likewise, if we wanted a function for the volume of a cube with each side having some length L, you may
recall volume of a rectangular box can be found by multiplying length by width by height, which are all
equal for a cube, giving the formula:

V ( L)  L  L  L  L3

These two functions are examples of power functions, functions that are some power of the variable.

Power Function

A power function is a function that can be represented in the form. Where the base is a variable and the
exponent, p, is a number

The constant and identity functions are power functions, since they can be written as f ( x)  x 0 and
f ( x)  x1 respectively.

The quadratic and cubic functions are both power functions with whole number powers: f ( x)  x 2 and
f ( x)  x 3 .

The reciprocal and reciprocal squared functions are both power functions with negative whole number
powers since they can be written as f ( x)  x 1 and f ( x)  x 2 .

The square and cube root functions are both power functions with fractional powers since they can be
written as f ( x)  x1 2 or f ( x)  x1 3 .
Characteristics of Power Functions

Shown above are the graphs of f ( x)  x 2 , f ( x)  x 4 , and f ( x)  x 6 , all even whole number powers. Notice
that all these graphs have a fairly similar shape, very similar to the quadratic toolkit, but as the power increases
the graphs flatten somewhat near the origin, and become steeper away from the origin.

To describe the behavior as numbers become larger and larger, we use the idea of infinity. The symbol for
positive infinity is  , and   for negative infinity. When we say that “x approaches infinity”, which can be
symbolically written as x   , we are describing a behavior – we are saying that x is getting large in the positive
direction.

With the even power function, as the input becomes large in either the positive or negative direction, the output
values become very large positive numbers. Equivalently, we could describe this by saying that as x approaches
positive or negative infinity, the f(x) values approach positive infinity. In symbolic form, we could write: as
x   , f (x)   .

f ( x)  x 7 f ( x)  x 5
Shown here are the graphs of
f ( x)  x , f ( x)  x , and f ( x)  x , all odd whole number
3 5 7
f ( x)  x 3
powers. Notice all these graphs look similar to the cubic toolkit,
but again as the power increases the graphs flatten near the
origin and become steeper away from the origin.

For these odd power functions, as x approaches negative infinity,


f(x) approaches negative infinity. As x approaches positive
infinity, f(x) approaches positive infinity. In symbolic form we
write: as x   , f (x)   and as x   , f (x)   .
Long Run Behaviour

The behavior of the graph of a function as the input takes on large negative values ( ) and large positive values ( )
as is referred to as the long run behavior of the function.

Polynomials

An oil pipeline bursts in the Gulf of Mexico, causing an oil slick in a roughly circular shape. The slick is currently 24
miles in radius, but that radius is increasing by 8 miles each week. If we wanted to write a formula for the area
covered by the oil slick, we could do so by composing two functions together. The first is a formula for the radius,
r, of the spill, which depends on the number of weeks, w, that have passed. Hopefully you recognized that this
relationship is linear:

r ( w)  24  8w

We can combine this with the formula for the area, A, of a circle:
A(r )  r 2

Composing these functions gives a formula for the area in terms of weeks:
A(w)  A(r (w))  A(24  8w)   (24  8w) 2

Multiplying this out gives the formula


A(w)  576  384w  64w 2

This formula is an example of a polynomial. A polynomial is simply the sum of terms each consisting of a
transformed power function with positive whole number power.

Terminology of Polynomial Functions


A polynomial is function that can be written as f ( x)  a 0  a1 x  a 2 x 2    a n x n

Each of the ai constants are called coefficients and can be positive, negative, or zero, and be whole numbers,
decimals, or fractions.

A term of the polynomial is any one piece of the sum that is any ai x i . Each individual term is a transformed
power function.

The degree of the polynomial is the highest power of the variable that occurs in the polynomial.

The leading term is the term containing the highest power of the variable: the term with the highest degree.

The leading coefficient is the coefficient of the leading term.

Because of the definition of the “leading” term we often rearrange polynomials so that the powers are
descending.
f ( x)  an x n  .....  a2 x 2  a1 x  a0
Long Run Behavior of Polynomials
For any polynomial, the long run behavior of the polynomial will match the long run behavior of the leading term.

Example 2.1
What can we determine about the long run behavior and degree of the equation for the
polynomial graphed here?

Since the output grows large and positive as the inputs grow large and positive, we describe the
long run behavior symbolically by writing: as x   , f (x)   . Similarly, as x   ,
f (x)   .

In words, we could say that as x values approach infinity, the function values approach infinity,
and as x values approach negative infinity the function values approach negative infinity.

We can tell this graph has the shape of an odd degree power function which has not been
reflected, so the degree of the polynomial creating this graph must be odd, and the leading
coefficient would be positive.

Short Run Behavior


Characteristics of the graph such as vertical and horizontal intercepts and the places the graph changes direction
are part of the short run behavior of the polynomial.

Like with all functions, the vertical intercept is where the graph crosses the vertical axis, and occurs when the
input value is zero. Since a polynomial is a function, there can only be one vertical intercept, which occurs at the
point (0, a0 ) . The horizontal intercepts occur at the input values that correspond with an output value of zero. It
is possible to have more than one horizontal intercept.

Notice that the polynomial in the previous example, which would be degree three if multiplied out, had three
horizontal intercepts and two turning points – places where the graph changes direction. We will now make a
general statement without justifying it – the reasons will become clear later in this chapter.

Intercepts and Turning Points of Polynomials


A polynomial of degree n will have:
At most n horizontal intercepts. An odd degree polynomial will always have at least one.
At most n-1 turning points
Example 2.2
What can we conclude about the graph of the polynomial shown here?

Based on the long run behavior, with the graph becoming large positive on both ends of the
graph, we can determine that this is the graph of an even degree polynomial. The graph has 2
horizontal intercepts, suggesting a degree of 2 or greater, and 3 turning points, suggesting a
degree of 4 or greater. Based on this, it would be reasonable to conclude that the degree is even
and at least 4, so it is probably a fourth degree polynomial.

2.2 QUADRATIC FUNCTIONS

In this section, we will explore the family of 2nd degree polynomials, the quadratic functions. While they share
many characteristics of polynomials in general, the calculations involved in working with quadratics is typically a
little simpler, which makes them a good place to start our exploration of short run behavior. In addition,
quadratics commonly arise from problems involving area and projectile motion, providing some interesting
applications.

Example 2.3
A backyard farmer wants to enclose a rectangular space for a new garden. She has purchased 80
feet of wire fencing to enclose 3 sides, and will put the 4th side against the backyard fence. Find a
formula for the area enclosed by the fence if the sides of fencing perpendicular to the existing
fence have length L.

In a scenario like this involving geometry, it is often helpful to


draw a picture. It might also be helpful to introduce a Garden L
temporary variable, W, to represent the side of fencing parallel
to the 4th side or backyard fence.
W
Since we know we only have 80 feet of fence available, we
know that L  W  L  80 , or more simply, 2L W  80 . This Backyard
allows us to represent the width, W, in terms of
L: W  80  2L

Now we are ready to write an equation for the area the fence encloses. We know the area of a
rectangle is length multiplied by width, so
A  LW  L(80  2 L)
A( L)  80 L  2 L2
This formula represents the area of the fence in terms of the variable length L.
Short run Behavior: Vertex

We now explore the interesting features of the graphs of quadratics. In addition to intercepts, quadratics have an
interesting feature where they change direction, called the vertex. You probably noticed that all quadratics are
related to transformations of the basic quadratic function f ( x)  x 2 .

Example 2.4
Write an equation for the quadratic graphed below as a transformation of f ( x)  x 2 , then
expand the formula and simplify terms to write the equation in standard polynomial form.

We can see the graph is the basic quadratic shifted to the left 2 and down 3, giving a formula in
the form g ( x)  a( x  2) 2  3 . By plugging in a point that falls on the grid, such as (0,-1), we
can solve for the stretch factor:
 1  a ( 0  2) 2  3
2  4a
1
a
2

1
Written as a transformation, the equation for this formula is g ( x)  ( x  2) 2  3 . To write
2
this in standard polynomial form, we can expand the formula and simplify terms:
1
g ( x)  ( x  2) 2  3
2
1
g ( x)  ( x  2)( x  2)  3
2
1 2
g ( x)  ( x  4 x  4)  3
2
1 2
g ( x)  x  2x  2  3
2
1 2
g ( x)  x  2x  1
2

Notice that the horizontal and vertical shifts of the basic quadratic determine the location of the vertex of the
parabola; the vertex is unaffected by stretches and compressions.
Check your Progress 1

A coordinate grid has been superimposed over the quadratic path of a basketball. Find an equation for
the path of the ball. Does he make the basket? (From http://blog.mrmeyer.com/?p=4778, © Dan Meyer,
CC-BY)

Forms of Quadratic Functions


The standard form of a quadratic function is f ( x)  ax 2  bx  c
The transformation form of a quadratic function is f ( x)  a( x  h) 2  k
The vertex of the quadratic function is located at (h, k), where h and k are the numbers in the transformation form
of the function. Because the vertex appears in the transformation form, it is often called the vertex form.

In the previous example, we saw that it is possible to rewrite a quadratic function given in transformation form and
rewrite it in standard form by expanding the formula. It would be useful to reverse this process, since the
transformation form reveals the vertex.

Expanding out the general transformation form of a quadratic gives:


f ( x)  a( x  h) 2  k  a( x  h)( x  h)  k
f ( x)  a( x 2  2 xh  h 2 )  k  ax 2  2ahx  ah 2  k

This should be equal to the standard form of the quadratic:


ax 2  2ahx  ah 2  k  ax 2  bx  c

The second degree terms are already equal. For the linear terms to be equal, the coefficients must be equal:
b
 2ah  b , so h  
2a
This provides us a method to determine the horizontal shift of the quadratic from the standard form. We could
likewise set the constant terms equal to find:
2
 b  b2 b2
ah  k  c , so k  c  ah  c  a 
2 2
 ca 2 c
 2a  4a 4a
In practice, though, it is usually easier to remember that k is the output value of the function when the input is h,
so k  f (h) .
Finding the Vertex of a Quadratic
For a quadratic given in standard form, the vertex (h, k) is located at:
b  b 
h , k  f ( h)  f  
2a  2a 

Example 2.5
Find the vertex of the quadratic f ( x)  2 x 2  6 x  7 . Rewrite the quadratic into
transformation form (vertex form).
b 6 6 3
The horizontal coordinate of the vertex will be at h     
2a 2(2) 4 2
2
3 3 3 5
The vertical coordinate of the vertex will be at f    2   6   7 
2 2 2 2
Rewriting into transformation form, the stretch factor will be the same as the a in the original
quadratic. Using the vertex to determine the shifts,
2
 3 5
f ( x)  2 x   
 2 2

Short run Behavior: Intercepts

As with any function, we can find the vertical intercepts of a quadratic by evaluating the function at an input of
zero, and we can find the horizontal intercepts by solving for when the output will be zero. Notice that depending
upon the location of the graph, we might have zero, one, or two horizontal intercepts.

zero horizontal intercepts one horizontal intercept two horizontal intercepts

Quadratic Formula
For a quadratic function given in standard form f ( x)  ax 2  bx  c , the quadratic formula gives the horizontal
intercepts of the graph of this function.
 b  b 2  4ac
x
2a

Example 2.6
A ball is thrown upwards from the top of a 40 foot high building at a speed of 80 feet per second. The
ball’s height above ground can be modeled by the equation H (t )  16t 2  80t  40 .
What is the maximum height of the ball?
When does the ball hit the ground?

To find the maximum height of the ball, we would need to know the vertex of the quadratic.
2
80 80 5 5 5 5
h   , k  H    16    80    40  140
2(16) 32 2 2 2 2

The ball reaches a maximum height of 140 feet after 2.5 seconds.

To find when the ball hits the ground, we need to determine when the height is zero – when H(t)
= 0. While we could do this using the transformation form of the quadratic, we can also use the
quadratic formula:
 80  80 2  4(16)( 40)  80  8960
t 
2(16)  32

Since the square root does not simplify nicely, we can use a calculator to approximate the values
of the solutions:
 80  8960  80  8960
t  5.458 or t  0.458
 32  32

The second answer is outside the reasonable domain of our model, so we conclude the ball will
hit the ground after about 5.458 seconds.

2.3 GRAPHS OF POLYNOMIAL FUNCTIONS

In the previous section we explored the short run behavior of quadratics, a special case of polynomials. In this
section we will explore the short run behavior of polynomials in general.

Short run Behavior: Intercepts

As with any function, the vertical intercept can be found by evaluating the function at an input of zero. Since this
is evaluation, it is relatively easy to do it for a polynomial of any degree.

To find horizontal intercepts, we need to solve for when the output will be zero. For general polynomials, this can
be a challenging prospect. While quadratics can be solved using the relatively simple quadratic formula, the
corresponding formulas for cubic and 4th degree polynomials are not simple enough to remember, and formulas
do not exist for general higher-degree polynomials. Consequently, we will limit ourselves to three cases:

1) The polynomial can be factored using known methods: greatest common factor and trinomial factoring.
2) The polynomial is given in factored form.
3) Technology is used to determine the intercepts.
Example 2.7
Find the horizontal intercepts of f ( x)  x 6  3x 4  2 x 2 .

We can attempt to factor this polynomial to find solutions for f(x) = 0.


x 6  3x 4  2 x 2  0 Factoring out the greatest common factor
2 4

x x  3x  2  0
2
 Factoring the inside as a quadratic in x2
x 2
x 2
 1x  2  0
2
Then break apart to find solutions

x 2

1  0 x 2

2 0
x 02
or x2  1 or x2  2
x0
x  1 x 2

This gives us 5 horizontal intercepts.

Example 2.8
Find the horizontal intercepts of h(t )  t 3  4t 2  t  6

Since this polynomial is not in factored form, has no


common factors, and does not appear to be factorable
using techniques we know, we can turn to technology to
find the intercepts.

Graphing this function, it appears there are horizontal


intercepts at t = -3, -2, and 1.

We could check these are correct by plugging in these


values for t and verifying that h(3)  h(2)  h(1)  0 .

Activity 2
1. Find the vertical and horizontal intercepts of the function f (t )  t 4  4t 2 .

Graphical Behavior at Intercepts

If we graph the function f ( x)  ( x  3)( x  2) 2 ( x  1) 3 , notice


that the behavior at each of the horizontal intercepts is different.

At the horizontal intercept x = -3, coming from the ( x  3) factor


of the polynomial, the graph passes directly through the
horizontal intercept. The factor is linear (has a power of 1), so
the behavior near the intercept is like that of a line - it passes
directly through the intercept. We call this a single zero, since
the zero corresponds to a single factor of the function.
At the horizontal intercept x = 2, coming from the ( x  2) 2 factor of the polynomial, the graph touches the axis
at the intercept and changes direction. The factor is quadratic (degree 2), so the behavior near the intercept is
like that of a quadratic – it bounces off of the horizontal axis at the intercept. Since ( x  2) 2  ( x  2)( x  2) ,
the factor is repeated twice, so we call this a double zero.

At the horizontal intercept x = -1, coming from the ( x  1) 3 factor of the polynomial, the graph passes through
the axis at the intercept, but flattens out a bit first. This factor is cubic (degree 3), so the behavior near the
intercept is like that of a cubic, with the same “S” type shape near the intercept that the toolkit x 3 has. We call
this a triple zero.

By utilizing these behaviors, we can sketch a reasonable graph of a factored polynomial function without
needing technology.

Graphical Behavior of Polynomials at Horizontal Intercepts


If a polynomial contains a factor of the form ( x  h) p , the behavior near the horizontal intercept h is
determined by the power on the factor.

p=1 p=2 p=3

Single zero Double zero Triple zero

For higher even powers 4,6,8 etc.… the graph will still bounce off of the horizontal axis but the graph will appear
flatter with each increasing even power as it approaches and leaves the axis.

For higher odd powers, 5,7,9 etc… the graph will still pass through the horizontal axis but the graph will appear
flatter with each increasing odd power as it approaches and leaves the axis.

Example 2.9
Sketch a graph of f ( x)  2( x  3) 2 ( x  5) .
This graph has two horizontal intercepts. At x = -3, the factor is squared, indicating the graph will
bounce at this horizontal intercept. At x = 5, the factor is not squared, indicating the graph will
pass through the axis at this intercept.
Additionally, we can see the leading term, if this polynomial were multiplied out, would be
 2x 3 , so the long-run behavior is that of a vertically reflected cubic, with the outputs
decreasing as the inputs get large positive, and the inputs increasing as the inputs get large
negative.
To sketch this we consider the following:
As x   the function f (x)   so we know the graph starts in the 2nd quadrant and is
decreasing toward the horizontal axis.

At (-3, 0) the graph bounces off of the horizontal axis and so the function must start increasing.

At (0, 90) the graph crosses the vertical axis at the vertical intercept.

Somewhere after this point, the graph must turn


back down or start decreasing toward the horizontal
axis since the graph passes through the next
intercept at (5,0).

As x   the function f (x)   so we know the


graph continues to decrease and we can stop
drawing the graph in the 4th quadrant.

Using technology we can verify that the resulting


graph will look like:

Solving Polynomial Inequalities

One application of our ability to find intercepts and sketch a graph of polynomials is the ability to solve
polynomial inequalities. It is a very common question to ask when a function will be positive and negative. We can
solve polynomial inequalities by either utilizing the graph, or by using test values.

Example 2.10
Solve ( x  3)( x  1) 2 ( x  4)  0

As with all inequalities, we start by solving the equality ( x  3)( x  1) 2 ( x  4)  0 , which has
solutions at x = -3, -1, and 4. We know the function can only change from positive to negative at
these values, so these divide the inputs into 4 intervals.

We could choose a testvalue in each interval and evaluate the function


f ( x)  ( x  3)( x  1) ( x  4) at each test value to determine if the function is positive or
2

negative in that interval

Interval Test x in interval f(test value) >0 or <0?


x < − 3 − 4 72 >0
− 3 < x < − 1 − 2 − 6 <0
− 1 < x < 4 0 − 12 <0
x>4 5 288 >0
On a number line this would look like:
From our test values, we can determine this function is positive when x < -3 or x > 4, or in
interval notation, (,3)  (4, )

We could have also determined on which intervals the function was positive by sketching a graph of the function.
We illustrate that technique in the next example

Activity 3
Given the function g ( x)  x 3  x 2  6 x use the methods that we have learned so far to find the vertical &
horizontal intercepts, determine where the function is negative and positive, describe the long run behavior
and sketch the graph without technology.

Writing Equations using Intercepts


Since a polynomial function written in factored form will have a horizontal intercept where each factor is equal
to zero, we can form a function that will pass through a set of horizontal intercepts by introducing a
corresponding set of factors.

Factored Form of Polynomials


If a polynomial has horizontal intercepts at x  x1 , x2 ,, xn , then the polynomial can be written in the factored
form
f ( x)  a( x  x1 ) p1 ( x  x 2 ) p2  ( x  x n ) pn
where the powers pi on each factor can be determined by the behavior of the graph at the corresponding
intercept, and the stretch factor a can be determined given a value of the function other than the horizontal
intercept.
Example 2.11
Write a formula for the polynomial function graphed here.

This graph has three horizontal intercepts: x = − 3, 2, and 5. At x = -3 and 5 the graph passes
through the axis, suggesting the corresponding factors of the polynomial will be linear. At x = 2
the graph bounces at the intercept, suggesting the corresponding factor of the polynomial will
be 2nd degree (quadratic). Together, this gives us:
f ( x)  a( x  3)( x  2) 2 ( x  5)

To determine the stretch factor, we can utilize another point on the graph. Here, the vertical
intercept appears to be (0,-2), so we can plug in those values to solve for a:

 2  a(0  3)(0  2) 2 (0  5)
 2  60a
1
a
30

1
The graphed polynomial appears to represent the function f ( x)  ( x  3)( x  2) 2 ( x  5) .
30

Activity 4
Given the graph, write a formula for the function shown.

Estimating Extrema

With quadratics, we were able to algebraically find the maximum or minimum value of the function by finding
the vertex. For general polynomials, finding these turning points is not possible without more advanced
techniques from calculus. Even then, finding where extrema occur can still be algebraically challenging. For now,
we will estimate the locations of turning points using technology to generate a graph.

Example 2.12
An open-top box is to be constructed by cutting out squares from w
each corner of a 14cm by 20cm sheet of plastic then folding up
the sides. Find the size of squares that should be cut out to w
maximize the volume enclosed by the box.

We will start this problem by drawing a picture, labeling the


width of the cut-out squares with a variable, w.
Notice that after a square is cut out from each end, it leaves a (14-2w) cm by (20-2w) cm
rectangle for the base of the box, and the box will be w cm tall. This gives the volume:
V (w)  (14  2w)(20  2w)w  280w  68w 2  4w3

Using technology to sketch a graph allows us to estimate the maximum value for the volume,
restricted to reasonable values for w: values from 0 to 7.

From this graph, we can estimate the maximum value is around 340, and occurs when the
squares are about 2.75cm square. To improve this estimate, we could use advanced features
of our technology, if available, or simply change our window to zoom in on our graph.

From this zoomed-in view, we can refine our estimate for the max volume to about 339, when
the squares are 2.7cm square.

2.4 RATIONAL FUNCTIONS


In the last few sections, we have built polynomials based on the positive whole number power functions. In this
section we explore functions based on power functions with negative integer powers, called rational functions.

Example 2.13
You plan to drive 100 miles. Find a formula for the time the trip will take as a function of the
speed you drive.
You may recall that multiplying speed by time will give you distance. If we let t represent the
drive time in hours, and v represent the velocity (speed or rate) at which we drive, then
vt  distance . Since our distance is fixed at 100 miles, vt  100 . Solving this relationship for
the time gives us the function we desired:
100
t (v )   100v 1
v

While this type of relationship can be written using the negative exponent, it is more common to see it written
as a fraction.

This particular example is one of an inversely proportional relationship – where one quantity is a constant
5
divided by the other quantity, like y  .
x
1
Notice that this is a transformation of the reciprocal toolkit function, f ( x ) 
x
Several natural phenomena, such as gravitational force and volume of sound, behave in a manner inversely
proportional to the square of another quantity. For example, the volume, V, of a sound heard at a distance d
k
from the source would be related by V  for some constant value k.
d2
1
These functions are transformations of the reciprocal squared toolkit function f ( x)  .
x2
We have seen the graphs of the basic reciprocal function and the squared reciprocal function from our study of
toolkit functions. These graphs have several important features.

1 1
f ( x)  f ( x) 
x x2

1
Let’s begin by looking at the reciprocal function, f ( x )  . As you well know, dividing by zero is not allowed
x
and therefore zero is not in the domain, and so the function is undefined at an input of zero.

Short run behavior:


As the input values approach zero from the left side (taking on very small, negative values), the function values
become very large in the negative direction (in other words, they approach negative infinity).

We write: as x  0  , f (x)   .

As we approach zero from the right side (small, positive input values), the function values become very large in
the positive direction (approaching infinity).

We write: as x  0  , f (x)   .

This behavior creates a vertical asymptote. An asymptote is a line that the graph approaches. In this case the
graph is approaching the vertical line x = 0 as the input becomes close to zero.
Long run behavior:
As the values of x approach infinity, the function values approach 0.
As the values of x approach negative infinity, the function values approach 0.
Symbolically: as x   , f ( x )  0
Based on this long run behavior and the graph we can see that the function approaches 0 but never actually
reaches 0, it just “levels off” as the inputs become large. This behavior creates a horizontal asymptote. In this
case the graph is approaching the horizontal line f ( x)  0 as the input becomes very large in the negative and
positive directions.

Vertical and Horizontal Asymptotes


A vertical asymptote of a graph is a vertical line x = a where the graph tends towards positive or negative
infinity as the inputs approach a. As x  a , f (x)   .

A horizontal asymptote of a graph is a horizontal line y  b where the graph approaches the line as the inputs
get large. As x   , f ( x )  b .

Example 2.14
Sketch a graph of the reciprocal function shifted two units to the left and up three units.
Identify the horizontal and vertical asymptotes of the graph, if any.

Transforming the graph left 2 and up 3 would result in the function


1
f ( x)   3 , or equivalently, by giving the terms a common denominator,
x2
3x  7
f ( x) 
x2

Shifting the toolkit function would give us


this graph. Notice that this equation is
undefined at x = -2, and the graph also is
showing a vertical asymptote at x = -2.
As x  2  , f ( x)   , and as
x  2  , f ( x )  

As the inputs grow large, the graph


appears to be leveling off at output values
of 3, indicating a horizontal asymptote at
y  3.
As x   , f ( x )  3 .
Notice that horizontal and vertical asymptotes get shifted left 2 and up 3 along with the
function.
In the previous example, we shifted a toolkit function in a way that resulted in a function of
3x  7
the form f ( x )  . This is an example of a more general rational function.
x2
Rational Function
A rational function is a function that can be written as the ratio of two polynomials, P(x) and Q(x).
P( x) a0  a1 x  a2 x   ap x p
2

f ( x)  
Q( x) b0  b1 x  b2 x 2   bq x q

Example 2.15
A large mixing tank currently contains 100 gallons of water, into which 5 pounds of
sugar have been mixed. A tap will open pouring 10 gallons per minute of water into
the tank at the same time sugar is poured into the tank at a rate of 1 pound per
minute. Find the concentration (pounds per gallon) of sugar in the tank after t minutes.

Notice that the amount of water in the tank is changing linearly, as is the amount of
sugar in the tank. We can write an equation independently for each:
water  100  10t
sugar  5  1t

The concentration, C, will be the ratio of pounds of sugar to gallons of water


5t
C (t ) 
100  10t

Finding Asymptotes and Intercepts


Given a rational function, as part of investigating the short run behavior we are interested in finding any vertical
and horizontal asymptotes, as well as finding any vertical or horizontal intercepts, as we have done in the past.
To find vertical asymptotes, we notice that the vertical asymptotes in our examples occur when the
denominator of the function is undefined. With one exception, a vertical asymptote will occur whenever the
denominator is undefined.

Example 2.16
x2
Find the vertical asymptotes of the function k ( x)  .
x2  4
To find the vertical asymptotes, we determine where this function will be undefined by
setting the denominator equal to zero:
x2  4  0
x2  4
x  2, 2
However, the numerator of this function is also equal
to zero when x = 2. Because of this, the function will
0
still be undefined at 2, since is undefined, but the
0
graph will not have a vertical asymptote at x = 2.
The graph of this function will have the vertical asymptote at x = -2, but at x = 2 the
graph will have a hole: a single point where the graph is not defined, indicated by an
open circle.
Vertical Asymptotes and Holes of Rational Functions
The vertical asymptotes of a rational function will occur where the denominator of the function is equal to zero
and the numerator is not zero.

A hole might occur in the graph of a rational function if an input causes both numerator and denominator to be
zero. In this case, factor the numerator and denominator and simplify; if the simplified expression still has a zero
in the denominator at the original input the original function has a vertical asymptote at the input, otherwise it
has a hole.

To find horizontal asymptotes, we are interested in the behavior of the function as the input grows large, so we
consider long run behavior of the numerator and denominator separately. Recall that a polynomial’s long run
behavior will mirror that of the leading term. Likewise, a rational function’s long run behavior will mirror that of
the ratio of the leading terms of the numerator and denominator functions.

There are three distinct outcomes when this analysis is done:

Case 1: The degree of the denominator > degree of the numerator


3x  2
Example: f ( x) 
x  4x  5
2

3x 3
In this case, the long run behavior is f ( x)   . This tells us that as the inputs grow large, this function will
x2 x
3
behave similarly to the function g ( x)  . As the inputs grow large, the outputs will approach zero, resulting in
x
a horizontal asymptote at y  0 .
As x   , f ( x )  0

Case 2: The degree of the denominator < degree of the numerator


3x 2  2
Example: f ( x) 
x5
3x 2
In this case, the long run behavior is f ( x)   3x . This tells us that as the inputs grow large, this function
x
will behave similarly to the function g ( x)  3x . As the inputs grow large, the outputs will grow and not level off,
so this graph has no horizontal asymptote.

As x   , f (x)   , respectively.

Ultimately, if the numerator is larger than the denominator, the long run behavior of the graph will mimic the
behavior of the reduced long run behavior fraction. As another example if we had the function
3x5  x 2 3x5
f ( x)  with long run behaviour f ( x)   3x 4 , the long run behavior of the graph would look
x3 x
similar to that of an even polynomial, and as x   , f (x)   .

Case 3: The degree of the denominator = degree of the numerator


3x 2  2
Example: f ( x)  2
x  4x  5
3x 2
In this case, the long run behavior is f ( x)   3 . This tells us that as the inputs grow large, this function
x2
will behave like the function g ( x)  3 , which is a horizontal line. As x   , f ( x )  3 , resulting in a
horizontal asymptote at y  3 .

Horizontal Asymptote of Rational Functions


The horizontal asymptote of a rational function can be determined by looking at the degrees of the numerator
and denominator.
Degree of denominator > degree of numerator: Horizontal asymptote at y  0
Degree of denominator < degree of numerator: No horizontal asymptote
Degree of denominator = degree of numerator: Horizontal asymptote at ratio of leading coefficients.

Example 2.17
Find the horizontal and vertical asymptotes of the function

( x  2)( x  3)
f ( x) 
( x  1)( x  2)( x  5)

First, note this function has no inputs that make both the numerator and denominator
zero, so there are no potential holes. The function will have vertical asymptotes when
the denominator is zero, causing the function to be undefined. The denominator will
be zero at x = 1, -2, and 5, indicating vertical asymptotes at these values.

The numerator has degree 2, while the denominator has degree 3. Since the degree of
the denominator is greater than the degree of the numerator, the denominator will
grow faster than the numerator, causing the outputs to tend towards zero as the
inputs get large, and so as x   , f ( x )  0 . This function will have a horizontal
asymptote at y  0 .

Intercepts

As with all functions, a rational function will have a vertical intercept when the input is zero, if the function is
defined at zero. It is possible for a rational function to not have a vertical intercept if the function is undefined at
zero.

Likewise, a rational function will have horizontal intercepts at the inputs that cause the output to be zero (unless
that input corresponds to a hole). It is possible there are no horizontal intercepts. Since a fraction is only equal
to zero when the numerator is zero, horizontal intercepts will occur when the numerator of the rational function
is equal to zero.
Since a rational function written in factored form will have a horizontal intercept where each factor of the
numerator is equal to zero, we can form a numerator that will pass through a set of horizontal intercepts by
introducing a corresponding set of factors. Likewise, since the function will have a vertical asymptote where each
factor of the denominator is equal to zero, we can form a denominator that will produce the vertical asymptotes
by introducing a corresponding set of factors.

Writing Rational Functions from Intercepts and Asymptotes


If a rational function has horizontal intercepts at x  x1 , x2 ,, xn , and vertical asymptotes at x  v1 , v2 ,, vm
then the function can be written in the form

( x  x1 ) p1 ( x  x2 ) p2  ( x  xn ) pn
f ( x)  a
( x  v1 ) q1 ( x  v 2 ) q2  ( x  vm ) qn

where the powers pi or qi on each factor can be determined by the behavior of the graph at the corresponding
intercept or asymptote, and the stretch factor a can be determined given a value of the function other than the
horizontal intercept, or by the horizontal asymptote if it is nonzero.

2.5 INVERSES AND RADICAL FUNCTIONS

In this section, we will explore the inverses of polynomial and rational functions, and in particular the radical
functions that arise in the process.

Example 2.18
A water runoff collector is built in the shape of a parabolic trough as shown below.
Find the surface area of the water in the trough as a function of the depth of the
water.

12 in
3ft

18 in

Since it will be helpful to have an equation for the parabolic cross-sectional shape, we
will impose a coordinate system at the cross section, with x measured horizontally
and y measured vertically, with the origin at the vertex of the parabola.
y

x
From this we find an equation for the parabolic shape. Since we placed the origin at
the vertex of the parabola, we know the equation will have form y ( x)  ax 2 . Our
equation will need to pass through the point (6,18), from which we can solve for the
stretch factor a:
18  a6 2
18 1
a 
36 2
1 2
Our parabolic cross section has equation y ( x)  x
2

Since we are interested in the surface area of the water, we are interested in
determining the width at the top of the water as a function of the water depth. For
any depth y the width will be given by 2x, so we need to solve the equation above for
x. However notice that the original function is not one-to-one, and indeed given any
output there are two inputs that produce the same output, one positive and one
negative.

To find an inverse, we can restrict our original function to a limited domain on which
it is one-to-one. In this case, it makes sense to restrict ourselves to positive x values.
On this domain, we can find an inverse by solving for the input variable:
1 2
y x
2
2y  x2
x   2y
This is not a function as written. Since we are limiting ourselves to positive x values,
we eliminate the negative solution, giving us the inverse function we’re looking for
x( y )  2 y

Since x measures from the center out, the entire width of the water at the top will be
2x. Since the trough is 3 feet (36 inches) long, the surface area will then be 36(2x), or
in terms of y:
Area  72 x  72 2 y
The previous example illustrated two important things:
1) When finding the inverse of a quadratic, we have to limit ourselves to a domain on which the function is
one-to-one.
2) The inverse of a quadratic function is a square root function. Both are toolkit functions and different types
of power functions.

Functions involving roots are often called radical functions.

Example 2.19
Find the inverse of f ( x)  ( x  2)2  3  x 2  4 x  1
From the transformation form of the function, we can see this is a transformed
quadratic with vertex at (2,-3) that opens upwards. Since the graph will be decreasing
on one side of the vertex, and increasing on the other side, we can restrict this
function to a domain on which it will be one-to-one by limiting the domain to x  2 .

To find the inverse, we will use the vertex form of the quadratic. We start by replacing
the f(x) with a simple variable y, then solve for x.
y  ( x  2) 2  3 Add 3 to both sides
y  3  ( x  2) 2 Take the square root
 y3x2 Add 2 to both sides
2 y 3  x

Of course, as written this is not a function. Since we restricted our original function to
a domain of x  2 , the outputs of the inverse should be the same, telling us to utilize
the + case:
1
x f ( y)  2  y  3

If the quadratic had not been given in vertex form, rewriting it into vertex form is
probably the best approach. Alternatively, we could have taken the standard equation
and rewritten it equal to zero:
0  x 2  4x  1  y

We would then be able to use the quadratic formula with a  1 , b  4 , and


c  (1  y ) , resulting in the same solutions we found above:
(4)  (4)2  4(1)(1  y) 12  4 y
x  2  2 3 y
2 2

Example 2.20
Find the inverse of the function f ( x)  5 x 3  1 .

This is a transformation of the basic cubic toolkit function, and based on our
knowledge of that function, we know it is one-to-one. Solving for the inverse by solving
for x

y  5x 3  1
y  1  5x 3
y 1
 x3
5
y 1
x  f 1 ( y )  3
5
Notice that this inverse is also a transformation of a power function with a fractional
power, x1/3.

Example 2.21
( x  2)( x  3)
Find the domain of the function f ( x)  .
( x  1)
Since a square root is only defined when the quantity under the radical is non-
( x  2)( x  3)
negative, we need to determine where  0 . A rational function can
( x  1)
change signs (change from positive to negative or vice versa) at horizontal intercepts
and at vertical asymptotes. For this equation, the graph could change signs at x = -2, 1,
and 3.

To determine on which intervals the rational expression is positive, we could evaluate


the expression at test values, or sketch a graph. While both approaches work equally
well, for this example we will use a graph.

This function has two horizontal intercepts, both of which exhibit linear behavior,
where the graph will pass through the intercept. There is one vertical asymptote,
corresponding to a linear factor, leading to a behavior similar to the basic reciprocal
toolkit function. There is a vertical intercept at (0, 6). This graph does not have a
horizontal asymptote, since the degree of the numerator is larger than the degree of
the denominator.

From the vertical intercept and horizontal intercept at x = -2, we can sketch the left
side of the graph. From the behavior at the asymptote, we can sketch the right side of
the graph.

From the graph, we can now tell on which


intervals this expression will be non-
negative, so the original function f(x) will be
defined.
f(x) has domain  2  x  1 or x  3 , or in
interval notation, [2,1)  [3, ) .

Like with finding inverses of quadratic


functions, it is sometimes desirable to find the inverse of a rational function,
particularly of rational functions that are the ratio of linear functions, such as our
concentration examples.
Self-Assessment Question
1. What point(s) do the toolkit power functions have in common?
2. Given the function determine the short run behavior.
3. Find the long run behavior of each function as x   and x  
1. f  x   x 4 2. f  x   x6 3. f  x   x3 4. f  x   x5
5. f  x    x2 6. f  x    x4 7. f  x    x7 8. f  x    x9
4. Find the degree and leading coefficient of each polynomial
1. 4x 7 2. 5x 6
3. 5  x 2 4. 6  3x  4 x3
5. 2 x 4  3x 2  x  1 6. 6 x5  2 x 4  x 2  3
7.  2 x  3 x  4 (3x  1) 8.  3x  1 x  1 (4x  3)

5. Find the long run behavior of each function as x   and x  


1. 2 x 4  3x 2  x  1 2. 6 x5  2 x 4  x 2  3
3. 3 x 2  x  2 4. 2 x3  x 2  x  3

6. What is the maximum number of x-intercepts and turning points for a polynomial of degree 5?
7. What is the maximum number of x-intercepts and turning points for a polynomial of degree 8?
8. What is the least possible degree of the polynomial function shown in each graph?

Answer to Check your Progress


Check your Progress 1
1. The path passes through the origin with vertex at (-4, 7).
7
h( x )   ( x  4) 2  7 . To make the shot, h(-7.5) would need to be
16
about 4. h(7.5)  1.64 ; he doesn’t make it.
2. g ( x)  x 2  6 x  13 in Standard form; g ( x)  ( x  3) 2  4 in
Transformation form
3. Vertical intercept at (0, 13), NO horizontal intercepts.
4. a. Vertex is a minimum value
b. Vertex is a maximum value
Attributes
David Lippman and Melonie Rasmussen, Open Text Bookstore, Precalculus: An
InvestigationofFunctions,”
Chapter 3: Polynomial and Rational Functions,” licensed under a CC BY-SA 3.0 license.

This chapter is part of Precalculus: An Investigation of Functions © Lippman & Rasmussen 2011.
This material is licensed under a Creative Commons CC-BY-SA license.
Unit 3
Limits
Structure
3.1 Introduction
3.2 A Preview of Calculus
3.3 The Limit of a Function
3.4 The Limit Laws
3.5 Continuity
3.6 The Precise Definition of a Limit
Summary
Keywords
Self-Assessment Questions

The text is adapted by Symbiosis Centre for Distance Learning under a Creative Commons
Attribution-NonCommercial-ShareAlike 4.0 International (CC BY-NC-SA 4.0) as requested by the
work’s creator or licensees. This license is available at https://creativecommons.org/licenses/by-nc-
sa/4.0/
Objectives
After going through this unit, you will be able to:
 Explain how the idea of a limit is involved in solving the tangent problem
 Recognize a tangent to a curve at a point as the limit of secant lines
 Explain how the idea of a limit is involved in solving the area problem
 Explain the relationship between one-sided and two-sided limits
 Recognize the basic limit laws
 Evaluate the limit of a function by factoring
 Explain the three conditions for continuity at a point
 State the theorem for limits of composite functions
 Describe the epsilon-delta definition of a limit

3.1 INTRODUCTION
Science fiction writers often imagine spaceships that can travel to far-off planets in distant galaxies. However,
back in 1905, Albert Einstein showed that a limit exists to how fast any object can travel. The problem is that
the faster an object moves, the more mass it attains (in the form of energy), according to the equation

where m0 is the object’s mass at rest, v is its speed, and c is the speed of light. What is this speed limit? (We
explore this problem further in Example 3.9.)
The idea of a limit is central to all of calculus. We begin this chapter by examining why limits are so important.
Then, we go on to describe how to find the limit of a function at a given point. Not all functions have limits at
all points, and we discuss what this means and how we can tell if a function does or does not have a limit at a
particular value. This chapter has been created in an informal, intuitive fashion, but this is not always enough if
we need to prove a mathematical statement involving limits. The last section of this chapter presents the more
precise definition of a limit and shows how to prove whether a function has a limit.
3.2 A PREVIEW OF CALCULUS
As we embark on our study of calculus, we shall see how its development arose from common solutions to
practical problems in areas such as engineering physics—like the space travel problem posed in the chapter
opener. Two key problems led to the initial formulation of calculus: (1) the tangent problem, or how to determine
the slope of a line tangent to a curve at a point; and (2) the area problem, or how to determine the area under a
curve.
The Tangent Problem and Differential Calculus
Rate of change is one of the most critical concepts in calculus. We begin our investigation of rates of change by

looking at the graphs of the three lines shown in Fig. 3.1.

Fig. 3.1 The rate of change of a linear function is constant in each of these three graphs, with the constant
determined by the slope
As we move from left to right along the graph of f(x) = −2x − 3, we see that the graph decreases at a constant rate.
For every 1 unit we move to the right along the x-axis, the y-coordinate decreases by 2 units. This rate of change
is determined by the slope (−2) of the line. Similarly, the slope of 1/2 in the function g(x) tells us that for every
change in x of 1 unit there is a corresponding change in y of 1/2 unit. The function h(x) = 2 has a slope of zero,
indicating that the values of the function remain constant. We see that the slope of each linear function indicates
the rate of change of the function.

Compare the graphs of these three functions with the graph of k(x) = x2 (Fig. 3.2). The graph of k(x) = x2 starts from
the left by decreasing rapidly, then begins to decrease more slowly and level off, and then finally begins to
increase—slowly at first, followed by an increasing rate of increase as it moves toward the right. Unlike a linear
function, no single number represents the rate of change for this function. We quite naturally ask: How do we
measure the rate of change of a nonlinear function?

Fig. 3.2 The function k(x) = x2 does not have a constant rate of change

We can approximate the rate of change of a function f(x) at a point (a, f(a)) on its graph by taking another point
(x, f(x)) on the graph of f(x), drawing a line through the two points, and calculating the slope of the resulting line.
⎝ ⎠
Such a line is called a secant line. Figure 3.3 shows a secant line to a function f(x) at a point (a, f(a)).

Fig. 3.3 The slope of a secant line through a point (a, f(a)) estimates the rate of change of the function at the
point (a, f(a)).
We formally define a secant line as follows:

Definition
The secant to the function f(x) through the points (a, f(a)) and (x, f(x)) is the line passing through these points.
Its slope is given by

(3.1)

The accuracy of approximating the rate of change of the function with a secant line depends on how close x is
to a. As we see in Fig. 3.4, if x is closer to a, the slope of the secant line is a better measure of the rate of change
of f (x) at a.

Fig. 3.4 As x gets closer to a, the slope of the secant line becomes a better approximation to the rate of change
of the function f(x) at a.

The secant lines themselves approach a line that is called the tangent to the function f(x) at a (Fig. 3.5). The
slope of the tangent line to the graph at a measures the rate of change of the function at a. This value also
represents the derivative of the function f(x) at a, or the rate of change of the function at a. This derivative is
denoted by f′(a). Differential calculus is the field of calculus concerned with the study of derivatives and their
applications.

Acitvity 1

For an interactive demonstration of the slope of a secant line that you can manipulate yourself, visit this
applet (Note: this site requires a Java browser plugin): Math Insight (http://www.openstaxcollege.org/l/
20_mathinsight) .
Fig. 3.5 Solving the Tangent Problem: As x approaches a, the secant lines approach the tangent line

Example 3.1
Illustrates how to find slopes of secant lines. These slopes estimate the slope of the tangent
line or, equivalently, the rate of change of the function at the point at which the slopes are
calculated.
Finding Slopes of Secant Lines
Estimate the slope of the tangent line (rate of change) to f(x) = x2 at x = 1 by finding slopes of
secant lines through (1, 1) and each of the following points on the graph of f(x) = x2.

Solution
Use the formula for the slope of a secant line from the definition.

The point in part b. is closer to the point (1, 1), so the slope of 2.5 is closer to the slope of the
tangent line. A good estimate for the slope of the tangent would be in the range of 2 to 2.5
(Fig. 3.6).
Fig. 3.6 The secant lines to f(x) = x2 at (1, 1) through (a) (2, 4) and (b) provide
successively closer approximations to the tangent line to f(x) = x2 at (1, 1).

We continue our investigation by exploring a related question. Keeping in mind that velocity may be thought of
as the rate of change of position, suppose that we have a function, s(t), that gives the position of an object along
a coordinate axis at any given time t. Can we use these same ideas to create a reasonable definition of the
instantaneous velocity at a given time t = a ? We start by approximating the instantaneous velocity with an average
velocity. First, recall that the speed of an object traveling at a constant rate is the ratio of the distance traveled to
the length of time it has traveled. We define the average velocity of an object over a time period to be the
change in its position divided by the length of the time period.

Definition
Let s(t) be the position of an object moving along a coordinate axis at time t. The average velocity
of the object over a time interval [a, t] where a < t (or [t, a] if t < a) is

(3.2)

As t is chosen closer to a, the average velocity becomes closer to the instantaneous velocity. Note that finding
the average velocity of a position function over a time interval is essentially the same as finding the slope of a
secant line to a function. Furthermore, to find the slope of a tangent line at a point a, we let the x-values approach
a in the slope of the secant line. Similarly, to find the instantaneous velocity at time a, we let the t-values
approach a in the average velocity. This process of letting x or t approach a in an expression is called taking a
limit. Thus, we may define the instantaneous velocity as follows.

Definition
For a position function s(t), the instantaneous velocity at a time t = a is the value that the average
velocities approach on intervals of the form [a, t] and [t, a] as the values of t become closer to a,
provided such a value exists.

Example 3.2
Illustrates this concept of limits and average velocity.
Finding Average Velocity
A rock is dropped from a height of 64 ft. It is determined that its height (in feet) above ground t
seconds later (for 0 ≤ t ≤ 2) is given by s(t) = −16t2 + 64. Find the average velocity of the rock over each
of the given time intervals. Use this information to guess the instantaneous velocity of the rock at time
t = 0.5.
a. [0.49, 0.5]
b. [0.5, 0.51]
Solution
Substitute the data into the formula for the definition of average velocity.

The instantaneous velocity is somewhere between −15.84 and −16.16 ft/sec. A good guess might be
−16 ft/sec.

The Area Problem and Integral Calculus


We now turn our attention to a classic question from calculus. Many quantities in physics—for example,
quantities of work—may be interpreted as the area under a curve. This leads us to ask the question: How can we
find the area between the graph of a function and the x-axis over an interval? (Fig. 3.7)

Fig. 3.7 The area problem: How do we find the area of the shaded region?
As in the answer to our previous questions on velocity, we first try to approximate the solution. We approximate
the area by dividing up the interval [a, b] into smaller intervals in the shape of rectangles. The approximation of
⎣ ⎦

the area comes from adding up the areas of these rectangles (Fig. 3.8).

Fig. 3.8 The area of the region under the curve is approximated by summing the areas of thin rectangles

As the widths of the rectangles become smaller (approach zero), the sums of the areas of the rectangles approach
the area between the graph of f(x) and the x-axis over the interval [a, b]. Once again, we find ourselves taking a
⎣ ⎦

limit. Limits of this type serve as a basis for the definition of the definite integral. Integral calculus is the study
of integrals and their applications.

Other Aspects of Calculus


So far, we have studied functions of one variable only. Such functions can be represented visually using graphs
in two dimensions; however, there is no good reason to restrict our investigation to two dimensions. Suppose,
for example, that instead of determining the velocity of an object moving along a coordinate axis, we want to
determine the velocity of a rock fired from a catapult at a given time, or of an airplane moving in three
dimensions. We might want to graph real-value functions of two variables or determine volumes of solids of the
type shown in Fig. 3.9. These are only a few of the types of questions that can be asked and answered using
multivariable calculus. Informally, multivariable calculus can be characterized as the study of the calculus of
functions of two or more variables. However, before exploring these and other ideas, we must first lay a
foundation for the study of calculus in one variable by exploring the concept of a limit.

Fig. 3.9 We can use multivariable calculus to find the volume between a surface defined by a function of two
variables and a plane

3.3 THE LIMIT OF A FUNCTION


The concept of a limit or limiting process, essential to the understanding of calculus, has been around for
thousands of years. In fact, early mathematicians used a limiting process to obtain better and better
approximations of areas of circles. Yet, the formal definition of a limit—as we know and understand it today—
did not appear until the late 19th century. We therefore begin our quest to understand limits, as our mathematical
ancestors did, by using an intuitive approach. At the end of this chapter, armed with a conceptual understanding
of limits, we examine the formal definition of a limit.
We begin our exploration of limits by taking a look at the graphs of the functions

which are shown in Fig. 3.10. In particular, let’s focus our attention on the behavior of each graph at and around
x = 2.

Fig. 3.10 These graphs show the behavior of three different functions around x = 2
Each of the three functions is undefined at x = 2, but if we make this statement and no other, we give a very
incomplete picture of how each function behaves in the vicinity of x = 2. To express the behavior of each graph
in the vicinity of 2 more completely, we need to introduce the concept of a limit.
Intuitive Definition of a Limit
Let’s first take a closer look at how the function f (x) = (x2 − 4)/(x − 2) behaves around x = 2 in Fig. 3.10. As the
values of x approach 2 from either side of 2, the values of y = f(x) approach 4. Mathematically, we say that the
limit of f(x) as x approaches 2 is 4. Symbolically, we express this limit as

From this very brief informal look at one limit, let’s start to develop an intuitive definition of the limit. We can
think of the limit of a function at a number a as being the one real number L that the functional values approach
as the x-values approach a, provided such a real number L exists. Stated more carefully, we have the following
definition:

Definition
Let f(x) be a function defined at all values in an open interval containing a, with the possible
exception of a itself, and let L be a real number. If all values of the function f(x) approach the
real number L as the values of x(≠ a) approach the number a, then we say that the limit of f(x) as
x approaches a is L. (More succinct, as x gets closer to a, f(x) gets closer and stays close to L.)
Symbolically, we express this idea as

(3.3)

We can estimate limits by constructing tables of functional values and by looking at their graphs. This process
is described in the following Problem-Solving Strategy.

We apply this Problem-Solving Strategy to compute a limit in Example 3.3.


Example 3.3
Evaluating a Limit Using a Table of Functional Values 1

Evaluate using a table of functional values.

Solution
We have calculated the values of f(x) = (sin x)/x for the values of x listed in Table 3.1.
Table 3.1 Table of Functional Values for

x x

−0.1 0.998334166468 0.1 0.998334166468

−0.01 0.999983333417 0.01 0.999983333417

−0.001 0.999999833333 0.001 0.999999833333

−0.0001 0.999999998333 0.0001 0.999999998333

Note: The values in this table were obtained using a calculator and using all the places
given in the calculator output.
As we read down each column, we see that the values in each column appear to be approaching

one. Thus, it is fairly reasonable to conclude that . A calculator-or computer-generated


graph of would be similar to that shown in Fig. 3.11, and it confirms our estimate.

Fig. 3.11 The graph of confirms the estimate from Table 3.1

At this point, we see from Example 3.3 that it may be just as easy, if not easier, to estimate a limit of a
function by inspecting its graph as it is to estimate the limit by using a table of functional values. In Example
3.4, we evaluate a limit exclusively by looking at a graph rather than by using a table of functional values.

Example 3.4
Evaluating a Limit Using a Graph

For g(x) shown in Fig. 3.12, evaluate

Fig. 3.12 The graph of g(x) includes one value not on a smooth curve

Solution
Despite the fact that g(−1) = 4, as the x-values approach −1 from either side, the g(x) values approach
3. Therefore, . Note that we can determine this limit without even knowing the algebraic
expression of the function.
Based on Example 3.4, we make the following observation: It is possible for the limit of a function to exist
at a point, and for the function to be defined at this point, but the limit of the function and the value of the
function at the point may be different.


2
Looking at a table of functional values or looking at the graph of a function provides us with useful insight
into the value of the limit of a function at a given point. However, these techniques rely too much on
guesswork. We eventually need to develop alternative methods of evaluating limits. These new methods
are more algebraic in nature and we explore them in the next section; however, at this point we introduce
two special limits that are foundational to the techniques to come.

Theorem 3.1: Two Important Limits


Let “a” be a real number and c be a constant.

(3.4)

(3.5)

We can make the following observations about these two limits.


I. For the first limit, observe that as x approaches a, so does f(x), because f(x) = x. Consequently,

II. For the second limit, consider Table 3.2.

Table 3.2 Table of Functional Values for

x f(x) = c x f(x) = c

a − 0.1 c a + 0.1 c

a − 0.01 c a + 0.01 c

a − 0.001 c a + 0.001 c

a − 0.0001 c a + 0.0001 c

Observe that for all values of x (regardless of whether they are approaching a), the values f(x) remain

constant at c. We have no choice but to conclude .

The Existence of a Limit


As we consider the limit in the next example, keep in mind that for the limit of a function to exist at a point,
the functional values must approach a single real-number value at that point. If the functional values do not
approach a single value, then the limit does not exist.
Example 3.5
Evaluating a Limit That Fails to Exist

Evaluate using a table of values.


Solution
Table 3.3 lists values for the function sin(1/x) for the given values of x.
Table 3.3 Table of Functional Values for

x sin(1/x) x sin(1/x)

−0.1 0.544021110889 0.1 −0.544021110889

−0.01 0.50636564111 0.01 −0.50636564111

−0.001 −0.8268795405312 0.001 0.826879540532

−0.0001 0.305614388888 0.0001 −0.305614388888

−0.00001 −0.035748797987 0.00001 0.035748797987

−0.000001 0.349993504187 0.000001 −0.349993504187

After examining the table of functional values, we can see that the y-values do not seem to approach
any one single value. It appears the limit does not exist. Before drawing this conclusion, let’s take a
more systematic approach. Take the following sequence of x-values approaching 0:

The corresponding y-values are


1, −1, 1, −1, 1, −1,….

At this point we can indeed conclude that does not exist. (Mathematicians frequently
abbreviate “does not exist” as DNE. Thus, we would write DNE.) The graph of
f(x) = sin(1/x) is shown in Fig. 3.13 and it gives a clearer picture of the behavior of sin(1/x) as x
approaches 0. You can see that sin(1/x) oscillates ever more wildly between −1 and 1 as x approaches
0.
Fig. 3.13 The graph of f (x) = sin(1/x) oscillates rapidly between −1 and 1 as x approaches 0

One-Sided Limits
Sometimes indicating that the limit of a function fails to exist at a point does not provide us with enough
information about the behavior of the function at that particular point. To see this, we now revisit the function
g(x) = |x − 2|/(x − 2) introduced at the beginning of the section (see Fig. 3.10(b)). As we pick values of x close

to 2, g(x) does not approach a single value, so the limit as x approaches 2 does not exist—that is,
DNE. However, this statement alone does not give us a complete picture of the behavior of the function
around the x-value 2. To provide a more accurate description, we introduce the idea of a one-sided limit.
For all values to the left of 2 (or the negative side of 2), g(x) = −1. Thus, as x approaches 2 from the left, g(x)
approaches −1. Mathematically, we say that the limit as x approaches 2 from the left is −1. Symbolically, we
express this idea as

Similarly, as x approaches 2 from the right (or from the positive side), g(x) approaches 1. Symbolically, we
express this idea as

We can now present an informal definition of one-sided limits.


Definition
We define two types of one-sided limits.
Limit from the left: Let f(x) be a function defined at all values in an open interval of the form z, and let L
be a real number. If the values of the function f(x) approach the real number L as the values of x (where
x < a) approach the number a, then we say that L is the limit of f (x) as x approaches a from the left.
Symbolically, we express this idea as

(3.6)
Limit from the right: Let f(x) be a function defined at all values in an open interval of the form (a, c), and
let L be a real number. If the values of the function f(x) approach the real number L as the values of x
(where x > a) approach the number a, then we say that L is the limit of f(x) as x approaches a from the
right. Symbolically, we express this idea as

(3.7)
Example 3.6
Evaluating One-Sided Limits

For the function , evaluate each of the following limits.

Solution
We can use tables of functional values again Table 3.4. Observe that for values of x less than 2, we
use f(x) = x + 1 and for values of x greater than 2, we use f(x) = x2 − 4.

Table 3.4 Table of Functional Values for

x f(x) = x + 1 x f(x) = x2 − 4
1.9 2.9 2.1 0.41
1.99 2.99 2.01 0.0401
1.999 2.999 2.001 0.004001
1.9999 2.9999 2.0001 0.00040001
1.99999 2.99999 2.00001 0.0000400001

Based on this table, we can conclude that a. and b. . Therefore, the


(two-sided) limit of f(x) does not exist at x = 2. Figure 3.14 shows a graph of f(x) and reinforces our
conclusion about these limits.

Fig. 3.14 The graph of has a break at x = 0

Let us now consider the relationship between the limit of a function at a point and the limits from the right and
left at that point. It seems clear that if the limit from the right and the limit from the left have a common value,
then that common value is the limit of the function at that point. Similarly, if the limit from the left and the limit
from the right take on different values, the limit of the function does not exist. These conclusions are summarized
in Relating One-Sided and Two- Sided Limits.

Theorem 3.2: Relating One-Sided and Two-Sided Limits


Let f (x) be a function defined at all values in an open interval containing a, with the possible exception of a
itself, and let L be a real number. Then,

Infinite Limits
Evaluating the limit of a function at a point or evaluating the limit of a function from the right and left at a point
helps us to characterize the behaviour of a function around a given value. As we shall see, we can also describe
the behavior of functions that do not have finite limits.
We now turn our attention to h(x) = 1/(x − 2)2, the third and final function introduced at the beginning of this
section (see Fig. 3.10(c)). From its graph we see that as the values of x approach 2, the values of h(x) = 1/(x − 2)2
become larger and larger and, in fact, become infinite. Mathematically, we say that the limit of h(x) as x
approaches 2 is positive infinity. Symbolically, we express this idea as


More generally, we define infinite limits as follows:
2

Definition
We define three types of infinite limits.
Infinite limits from the left: Let f(x) be a function defined at all values in an open interval of the form (b, a).
i. If the values of f (x) increase without bound as the values of x (where x < a) approach the number a, then
we say that the limit as x approaches a from the left is positive infinity and we write

(3.8)
ii. If the values of f (x) decrease without bound as the values of x (where x < a) approach the number a, then
we say that the limit as x approaches a from the left is negative infinity and we write

(3.9)
Infinite limits from the right: Let f (x) be a function defined at all values in an open interval of the form (a,
c).
i. If the values of f (x) increase without bound as the values of x (where x > a) approach the number a, then
we say that the limit as x approaches a from the left is positive infinity and we write

(3.10)
ii. If the values of f(x) decrease without bound as the values of x (where x > a) approach the number a, then
we say that the limit as x approaches a from the left is negative infinity and we write

(3.11)
Two-sided infinite limit: Let f (x) be defined for all x ≠ a in an open interval containing a.
i. If the values of f(x) increase without bound as the values of x (where x ≠ a) approach the number a, then
we say that the limit as x approaches a is positive infinity and we write

(3.12)
ii. If the values of f(x) decrease without bound as the values of x (where x ≠ a) approach the number a, then
we say that the limit as x approaches a is negative infinity and we write

(3.13)

Example 3.7
Recognizing an Infinite Limit
Evaluate each of the following limits, if possible. Use a table of functional values and graph f (x) =
1/x to confirm your conclusion.

Solution
Begin by constructing a table of functional values.

Table 3.5 Table of functional values for

x x

−0.1 −10 0.1 10


−0.01 −100 0.01 100
−0.001 −1000 0.001 1000
−0.0001 −10,000 0.0001 10,000
−0.00001 −100,000 0.00001 100,000

−0.000001 −1,000,000 0.000001 1,000,000

a. The values of 1/x decrease without bound as x approaches 0 from the left. We conclude that

b. The values of 1/x increase without bound as x approaches 0 from the right. We conclude that


c. Since and0 have different values, we conclude that
The graph of f(x) = 1/x in Fig. 3.15 confirms these conclusions.

Fig. 3.15 The graph of f (x) = 1/x confirms that the limit as x approaches 0 does not exist

It is useful to point out that functions of the form f(x) = 1/(x − a)n, where n is a positive integer, have infinite
limits as x approaches a from either the left or right (Fig. 3.16). These limits are summarized in Infinite Limits
from Positive Integers.

Fig. 3.16 The function f (x) = 1/(x − a)n has infinite limits at a
Theorem 3.3: Infinite Limits from Positive Integers
If n is a positive even integer, then

If n is a positive odd integer, then

And

We should also point out that in the graphs of f(x) = 1/(x − a)n, points on the graph having x-coordinates very near
to a are very close to the vertical line x = a. That is, as x approaches a, the points on the graph of f(x) are closer
to the line x = a. The line x = a is called a vertical asymptote of the graph. We formally define a vertical asymptote
as follows:

Definition
Let f(x) be a function. If any of the following conditions hold, then the line x = a is a vertical asymptote of
f(x).

In the next example we put our knowledge of various types of limits to use to analyze the behavior of a function
at several different points.
Example 3.8
Behavior of a Function at Different Points
Use the graph of f(x) in Fig. 3.17 to determine each of the following values:
Fig. 3.17 The graph shows f (x)
Solution
Using Infinite Limits from Positive Integers and the graph for reference, we arrive at the following
values:

Example 3.9
Chapter Opener: Einstein’s Equation

Fig. 3.18 (credit: NASA)

In the unit opener we mentioned briefly how Albert Einstein showed that a limit exists to how fast
any object can travel. Given Einstein’s equation for the mass of a moving object, what is the value
of this bound?

Solution
Our starting point is Einstein’s equation for the mass of a moving object,

where m0 is the object’s mass at rest, v is its speed, and c is the speed of light. To see how the mass
changes at high speeds, we can graph the ratio of masses m/m0 as a function of the ratio of speeds,
v/c (Fig. 3.19).

Fig. 3.19 This graph shows the ratio of masses as a function of the ratio of speeds in Einstein’s
equation for the mass of a moving object

We can see that as the ratio of speeds approaches 1—that is, as the speed of the object approaches
the speed of light—the ratio of masses increases without bound. In other words, the function has a
vertical asymptote at v/c = 1. We can try a few values of this ratio to test this idea.

Table 3.6 Ratio of Masses and Speeds for a Moving Object

0.99 0.1411 7.089

0.999 0.0447 22.37

0.9999 0.0141 70.71

Thus, according to Table 3.6, if an object with mass 100 kg is traveling at 0.9999c, its mass becomes
7071 kg. Since no object can have an infinite mass, we conclude that no object can travel at or more
than the speed of light.

3.4 THE LIMIT LAWS


In the previous section, we evaluated limits by looking at graphs or by constructing a table of values. In this
section, we establish laws for calculating limits and learn how to apply these laws. In the Student Project at the
end of this section, you have the opportunity to apply these limit laws to derive the formula for the area of a circle
by adapting a method devised by the Greek mathematician Archimedes. We begin by restating two useful limit
results from the previous section. These two results, together with the limit laws, serve as a foundation for
calculating many limits.
Evaluating Limits with the Limit Laws
The first two limit laws were stated in Two Important Limits and we repeat them here. These basic results,
together with the other limit laws, allow us to evaluate limits of many algebraic functions.

Theorem 3.4: Basic Limit Results


For any real number a and any constant c,

(3.14)
(3.15)

Example 3.10
Evaluating a Basic Limit
Evaluate each of the following limits using Basic Limit Results.

Solution

a. The limit of x as x approaches a is a:

b. The limit of a constant is that constant:

We now take a look at the limit laws, the individual properties of limits. The proofs that these laws hold are
omitted here.

Theorem 3.5: Limit Laws

We now practice applying these limit laws to evaluate a limit.


Example 3.11
Evaluating a Limit Using Limit Laws
Use the limit laws to evaluate .
Solution
Let’s apply the limit laws one step at a time to be sure we understand how they work. We need to
keep in mind the requirement that, at each application of a limit law, the new limits must exist for the
limit law to be applied.
Example 3.12
Using Limit Laws Repeatedly

Use the limit laws to evaluate


Solution
To find this limit, we need to apply the limit laws several times. Again, we need to keep in mind that
as we rewrite the limit in terms of other limits, each new limit must exist for the limit law to be
applied.

Activity 1

Use the limit laws to evaluate . In each step, indicate the limit law applied.

Limits of Polynomial and Rational Functions


By now you have probably noticed that, in each of the previous examples, it has been the case that
. This is not always true, but it does hold for all polynomials for any choice of a and for all
rational functions at all values of a for which the rational function is defined.

Theorem 3.6: Limits of Polynomial and Rational Functions


Let p(x) and q(x) be polynomial functions. Let a be a real number. Then,

To see that this theorem holds, consider the polynomial . By


applying the sum, constant multiple, and power laws, we end up with

It now follows from the quotient law that if p(x) and q(x) are polynomials for which q(a) ≠ 0, then

Example 3.13 applies this result.

Example 3.13
Evaluating a Limit of a Rational Function

Evaluate the .

Solution

Since 3 is in the domain of the rational function , we can calculate the limit by
substituting 3 for x into the function. Thus,

Additional Limit Evaluation Techniques


As we have seen, we may evaluate easily the limits of polynomials and limits of some (but not all) rational
functions by direct substitution. However, as we saw in the introductory section on limits, it is certainly
possible for to exist when f(a) is undefined. The following observation allows us to evaluate many
limits of this type: If for all over some open interval containing a, then .

To understand this idea better, consider the limit .


The function

and the function g(x) = x + 1 are identical for all values of x ≠ 1. The graphs of these two functions are shown
in Fig. 3.20.
Fig. 3.20 The graphs of f (x) and g(x) are identical for all x ≠ 1. Their limits at 1 are equal

We see that

The limit has the form ( )gx , where and . (In this case, we say that f(x)/g(x)
has the indeterminate form 0/0.) The following Problem-Solving Strategy provides a general outline for
evaluating limits of this type.

Problem-Solving Strategy: Calculating a Limit When f(x)/g(x) has the Indeterminate Form 0/0

1. First, we need to make sure that our function has the appropriate form and cannot be evaluated
immediately using the limit laws.
2. We then need to find a function that is equal to h(x) = f(x)/g(x) for all x ≠ a over some interval
containing
a. To do this, we may need to try one or more of the following steps:
a. If f (x) and g(x) are polynomials, we should factor each function and cancel out any common
factors.
b. If the numerator or denominator contains a difference involving a square root, we should
try multiplying the numerator and denominator by the conjugate of the expression involving
the square root.
c. If f (x)/g(x) is a complex fraction, we begin by simplifying it.
Last, we apply the limit laws.
Example 3.14
Evaluating a Limit by Multiplying by a Conjugate

Evaluate .

Solution

Step 1 has the form 0/0 at −1. Let’s begin by multiplying by , the conjugate of
, on the numerator and denominator:

Step 2. We then multiply out the numerator. We don’t multiply out the denominator because we are
hoping that the (x + 1) in the denominator cancels out in the end:

Step 3. Then we cancel:

Step 4. Last, we apply the limit laws:

Example 3.15
Evaluating a Limit When the Limit Laws Do Not Apply

Evaluate .
Solution
Both 1/x and 5/x(x − 5) fail to have a limit at zero. Since neither of the two functions has a limit at
zero, we cannot apply the sum law for limits; we must use a different strategy. In this case, we find
the limit by performing addition and then applying one of our previous strategies. Observe that

Let’s now revisit one-sided limits. Simple modifications in the limit laws allow us to apply them to one-sided

limits. For example, to apply the limit laws to a limit of the form we require the function h(x) to

be defined over an open interval of the form (b, a); for a limit of the form , we require the function
h(x) to be defined over an open interval of the form (a, c).
Example 3.16
Evaluating a Two-Sided Limit Using the Limit Laws

For

Solution
Figure 3.21 illustrates the function f (x) and aids in our understanding of these limits.

Fig. 3.21 This graph shows a function f(x)

a. Since f(x) = 4x − 3 for all x in (−∞, 2), replace f(x) in the limit with 4x − 3 and apply the limit
laws:

b. Since f (x) = (x − 3)2 for all x in (2, +∞), replace f (x) in the limit with (x − 3)2 and apply the
limit laws:

c. Since and , we conclude that does not exist.

We now turn our attention to evaluating a limit of the form where where K ≠ 0 and
That is, f(x)/g(x) has the form K/0, K≠0 at a.

Example 3.17
Evaluating a Limit of the Form K/0, K ≠ 0 Using the Limit Laws

Solution
Step 1. After substituting in x = 2, we see that this limit has the form −1/0. That is, as x approaches
2 from the left, the numerator approaches −1; and the denominator approaches 0. Consequently, the

magnitude of becomes infinite. To get a better idea of what the limit is, we need to factor the
denominator:

Step 2. Since x − 2 is the only part of the denominator that is zero when 2 is substituted, we then
separate 1/(x − 2) from the rest of the function:

Step 3. and . Therefore, the product of (x–3)/x and 1/(x-2) has a limit
of +∞:

The Squeeze Theorem


The techniques we have developed thus far work very well for algebraic functions, but we are still unable to
evaluate limits of very basic trigonometric functions. The next theorem, called the squeeze theorem, proves
very useful for establishing basic trigonometric limits. This theorem allows us to calculate limits by “squeezing”
a function, with a limit at a point a that is unknown, between two functions having a common known limit at a.
Fig. 3.22 illustrates this idea.

Fig. 3.22 The Squeeze Theorem applies when

Theorem 3.7: The Squeeze Theorem


Let f(x), g(x), and h(x) be defined for all x ≠ a over an open interval containing a. If
f(x) ≤ g(x) ≤ h(x)
for all x ≠ a in an open interval containing a and

where L is a real number, then .


Example 3.18
Applying the Squeeze Theorem

Apply the squeeze theorem to evaluate .


Solution

Because −1 ≤ cos x ≤ 1 for all x, we have − |x|≤ x cos x ≤ |x|. Since from

the squeeze theorem, we obtain . The graphs of f (x) = − |x|, g(x) = x cos x, and
h(x) = |x| are shown in Fig. 3.23.

Fig. 3.23 The graphs of f (x), g(x), and h(x) are shown around the point x = 0

We now use the squeeze theorem to tackle several very important limits. Although this discussion is
somewhat lengthy, these limits prove invaluable for the development
2 of the material in both the next section

and the next chapter. The first of these limits is . Consider the unit circle shown in Fig. 3.24. In
the figure, we see that sinθ is the y-coordinate on the unit circle and it corresponds to the line segment
shown in blue. The radian measure of angle θ is the length of the arc it subtends on the unit circle. Therefore,
we see that for .

Fig. 3.24 The sine function is shown as a line on the unit circle

Because and by using the squeeze theorem we conclude that


To see that as well, observe that for and hence,
Consequently, It follows that . An application of the squeeze theorem

produces the desired limit. Thus, since and

(3.16)

Next, using the identity for , we see that

(3.17)

We now take a look at a limit that plays an important role in later chapters—namely, . To evaluate this
limit, we use the unit circle in Fig. 3.25. Notice that this figure adds one additional triangle to Fig. 3.25. We see
that the length of the side opposite angle θ in this new triangle is tanθ. Thus, we see that for
.

Fig. 3.25 The sine and tangent functions are shown as lines on the unit circle

By dividing by sinθ in all parts of the inequality, we obtain

Equivalently, we have

Since , we conclude that . By applying a manipulation similar to that used in

demonstrating that we can show that

(3.18)

3.5 CONTINUITY
Many functions have the property that their graphs can be traced with a pencil without lifting the pencil from the
page. Such functions are called continuous. Other functions have points at which a break in the graph occurs,
but satisfy this property over intervals contained in their domains. They are continuous on these intervals and
are said to have a discontinuity at a point where a break occurs.
We begin our investigation of continuity by exploring what it means for a function to have continuity at a point.
Intuitively, a function is continuous at a particular point if there is no break in its graph at that point.
Continuity at a Point
Before we look at a formal definition of what it means for a function to be continuous at a point, let’s consider
various functions that fail to meet our intuitive notion of what it means to be continuous at a point. We then create
a list of conditions that prevent such failures.
Our first function of interest is shown in Fig. 3.26. We see that the graph of f(x) has a hole at a. In fact, f(a) is
undefined. At the very least, for f(x) to be continuous at a, we need the following condition:

Fig. 3.26 The function f(x) is not continuous at a because f(a) is undefined

However, as we see in Fig. 3.27, this condition alone is insufficient to guarantee continuity at the point a.
Although f(a) is defined, the function has a gap at a. In this example, the gap exists because does not
exist. We must add another condition for continuity at a—namely,

Fig. 3.27 The function f(x) is not continuous at a because does not exist

However, as we see in Fig. 3.28, these two conditions by themselves do not guarantee continuity at a point. The
function in this figure satisfies both of our first two conditions, but is still not continuous at a. We must add a
third condition to our list:

Fig. 3.28 The function f (x) is not continuous at a because


Now we put our list of conditions together and form a definition of continuity at a point.

Definition
A function f(x) is continuous at a point a if and only if the following three conditions are satisfied:

A function is discontinuous at a point a if it fails to be continuous at a.

The following procedure can be used to analyze the continuity of a function at a point using this definition.

Problem-Solving Strategy: Determining Continuity at a Point


1. Check to see if f (a) is defined. If f (a) is undefined, we need go no further. The function is not continuous
at a. If f (a) is defined, continue to step 2.

2. Compute . In some cases, we may need to do this by first computing . If

does not exist (that is, it is not a real number), then the function is not continuous at “a” and the
problem is solved. If exists, then continue to step 3.

3. Compare f(a) and . If , then the function is not continuous at a. If


, then the function is continuous at a.

The next three examples demonstrate how to apply this definition to determine whether a function is
continuous at a given point. These examples illustrate situations in which each of the conditions for continuity
in the definition succeed or fail.

Example 3.19
Determining Continuity at a Point, Condition 1
Using the definition, determine whether the function f (x) = (x2 − 4)/(x − 2) is continuous at x = 2.
Justify the conclusion.
Solution
Let’s begin by trying to calculate f(2). We can see that f(2) = 0/0, which is undefined. Therefore,
is discontinuous at 2 because f(2) is undefined. The graph of f(x) is shown in Fig. 3.29.

Fig. 3.29 The function f (x) is discontinuous at 2 because f(2) is undefined


Example 3.20
Determining Continuity at a Point, Condition 2

Using the definition, determine whether the function is continuous at x = 3.


Justify the conclusion.
Solution
Let’s begin by trying to calculate f(3).

Thus, f(3) is defined. Next, we calculate . To do this, we must compute and

Therefore, does not exist. Thus, f(x) is not continuous at 3. The graph of f(x) is shown in
Fig. 3.30.

Fig. 3.30 The function f(x) is not continuous at 3 because does not exist

Example 3.21
Determining Continuity at a Point, Condition 3

Using the definition, determine whether the function is continuous at x = 0.


Solution
First, observe that
f(0) = 1.
Next,

Last, compare f(0) and . We see that

Since all three of the conditions in the definition of continuity are satisfied, f(x) is continuous at x = 0.
By applying the definition of continuity and previously established theorems concerning the evaluation of limits,
we can state the following theorem.

Theorem 3.8: Continuity of Polynomials and Rational Functions


Polynomials and rational functions are continuous at every point in their domains.
Proof
Previously, we showed that if p(x) and q(x) are polynomials, for every polynomial p(x) and
as long as q(a) ≠ 0. Therefore, polynomials and rational functions are continuous on their
domains.
We now apply Continuity of Polynomials and Rational Functions to determine the points at which a given
rational function is continuous.

Example 3.22
Continuity of a Rational Function

For what values of x is continuous?

Solution
The rational function is continuous for every value of x except x = 5.

Types of Discontinuities
As we have seen in Example 3.19 and Example 3.20, discontinuities take on several different appearances. We
classify the types of discontinuities we have seen thus far as removable discontinuities, infinite discontinuities,
or jump discontinuities. Intuitively, a removable discontinuity is a discontinuity for which there is a hole in the
graph, a jump discontinuity is a non-infinite discontinuity for which the sections of the function do not meet
up, and an infinite discontinuity is a discontinuity located at a vertical asymptote. Figure 3.31 illustrates the
differences in these types of discontinuities. Although these terms provide a handy way of describing three
common types of discontinuities, keep in mind that not all discontinuities fit neatly into these categories.

Fig. 3.31 Discontinuities are classified as (a) removable, (b) jump, or (c) infinite

These three discontinuities are formally defined as follows:


Definition

Example 3.23
Classifying a Discontinuity

In Example 3.19, we showed that is discontinuous at x = 2. Classify this discontinuity


as removable, jump, or infinite.

Solution

Example 3.24
Classifying a Discontinuity

In Example 3.20, we showed that is discontinuous at x = 3. Classify this


discontinuity as removable, jump, or infinite.
Solution

Example 3.25
Classifying a Discontinuity
Determine whether is continuous at −1. If the function is discontinuous at −1, classify
the discontinuity as removable, jump, or infinite.
Solution
The function value f(−1) is undefined. Therefore, the function is not continuous at −1. To determine
the type of discontinuity, we must determine the limit at −1. We see that
Therefore, the function has an infinite discontinuity at −1.
Continuity over an Interval
Now that we have explored the concept of continuity at a point, we extend that idea to continuity over an
interval. As we develop this idea for different types of intervals, it may be useful to keep in mind the intuitive
idea that a function is continuous over an interval if we can use a pencil to trace the function between any two
points in the interval without lifting the pencil from the paper. In preparation for defining continuity on an interval,
we begin by looking at the definition of what it means for a function to be continuous from the right at a point
and continuous from the left at a point.

Continuity from the Right and from the Left


A function f(x) is said to be continuous from the right at a if

A function f (x) is said to be continuous from the left at a if

A function is continuous over an open interval if it is continuous at every point in the interval. A function f (x) is

continuous over a closed interval of the form a, b if it is continuous at every point in (a, b) and is continuous
⎡ ⎤

from the right at a and is continuous from the left at b. Analogously, a function f (x) is continuous over an
⎣ ⎦

interval of the form (a, b if it is continuous over (a, b) and is continuous from the left at b. Continuity over other

types of intervals are defined in a similar fashion.


Requiring that and ensures that we can trace the graph of the function from
the point (a, f(a)) to the point (b, f(b)) without lifting the pencil . If, for example, , we would need
to lift our pencil to jump from f(a) to the graph of the rest of the function over (a, b].

Example 3.26
Continuity on an Interval
State the interval(s) over which the function is continuous.
Solution
Since is a rational function, it is continuous at every point in its domain. The domain of f(x)
is the set (−∞, −2) ∪ (−2, 0) ∪ (0, +∞). Thus, f(x) is continuous over each of the intervals (−∞, −2), (−2, 0), and
(0, +∞).

Example 3.27
Continuity over an Interval
State the interval(s) over which the function is continuous.
Solution

The Composite Function Theorem allows us to expand our ability to compute limits. In particular, this theorem
ultimately allows us to demonstrate that trigonometric functions are continuous over their domains.

Theorem 3.9: Composite Function Theorem


Before we move on to Example 3.28, recall that earlier, in the section on limit laws, we showed

. Consequently, we know that f (x) = cos x is continuous at 0. In Example 3.28 we see


how to combine this result with the composite function theorem.

Example 3.28
Limit of a Composite Cosine Function

Evaluate
Solution

The proof of the next theorem uses the composite function theorem as well as the continuity of f(x) = sin x and
g(x) = cos x at the point 0 to show that trigonometric functions are continuous over their entire domains.

Theorem 3.10: Continuity of Trigonometric Functions


Trigonometric functions are continuous over their entire domains.

Proof
We begin by demonstrating that cos x is continuous at every real number. To do this, we must show that
for all values of a.

The proof that sin x is continuous at every real number is analogous. Because the remaining trigonometric
functions may be expressed in terms of sin x and cos x, their continuity follows from the quotient limit law.

As you can see, the composite function theorem is invaluable in demonstrating the continuity of trigonometric
functions. As we continue our study of calculus, we revisit this theorem many times.

The Intermediate Value Theorem


Functions that are continuous over intervals of the form [a, b], where a and b are real numbers, exhibit many useful
properties. Throughout our study of calculus, we will encounter many powerful theorems concerning such
functions. The first of these theorems is the Intermediate Value Theorem.
Theorem 3.11: The Intermediate Value Theorem
Let f be continuous over a closed, bounded interval [a, b]. If z is any real number between f(a) and f(b), then
there is a number c in [a, b] satisfying f(c) = z in Fig. 3.32.

Fig. 3.32 There is a number c ∈ [a, b] that satisfies f(c) = z

Example 3.29
Application of the Intermediate Value Theorem

Show that f(x) = x – cos x has at least one zero.


Solution

Example 3.30
When Can You Apply the Intermediate Value Theorem?
If f(x) is continuous over [0, 2], f(0) > 0 and f(2) > 0, can we use the Intermediate Value Theorem to
conclude that f(x) has no zeros in the interval [0, 2]? Explain.

Solution
No. The Intermediate Value Theorem only allows us to conclude that we can find a value between
f(0) and f(2); it doesn’t allow us to conclude that we can’t find other values. To see this more clearly,
consider the function f(x) = (x − 1)2. It satisfies f (0) = 1 > 0, f(2) = 1 > 0, and f(1) = 0.

3.6 THE PRECISE DEFINITION OF A LIMIT


By now you have progressed from the very informal definition of a limit in the introduction of this chapter to
the intuitive understanding of a limit. At this point, you should have a very strong intuitive sense of what the
limit of a function means and how you can find it. In this section, we convert this intuitive idea of a limit into a
formal definition using precise mathematical language. The formal definition of a limit is quite possibly one of
the most challenging definitions you will encounter early in your study of calculus; however, it is well worth
any effort you make to reconcile it with your intuitive notion of a limit. Understanding this definition is the key
that opens the door to a better understanding of calculus.

Quantifying Closeness
Before stating the formal definition of a limit, we must introduce a few preliminary ideas. Recall that the distance
between two points a and b on a number line is given by |a − b|.
 The statement |f (x) − L| < ε may be interpreted as: The distance between f (x) and L is less than ε.
 The statement 0 < |x − a| < δ may be interpreted as: x ≠ a and the distance between x and a is less than δ.
It is also important to look at the following equivalences for absolute value:
 The statement |f (x) − L| < ε is equivalent to the statement L − ε < f (x) < L + ε.
 The statement 0 < |x − a| < δ is equivalent to the statement a − δ < x < a + δ and x ≠ a.
With these clarifications, we can state the formal epsilon-delta definition of the limit.

Definition

This definition may seem rather complex from a mathematical point of view, but it becomes easier to
understand if we break it down phrase by phrase. The statement itself involves something called a universal
quantifier (for every ε > 0), an existential quantifier (there exists a δ > 0), and, last, a conditional statement (if 0
< |x − a| < δ, then |f (x) − L|< ε). Let’s take a look at Table 3.7, which breaks down the definition and translates
each part.

Table 3.7 Translation of the Epsilon-Delta Definition of the Limit


We can get a better handle on this definition by looking at the definition geometrically. Figure 3.33 show
possible values of δ for various choices of ε > 0 for a given function f(x), a number a, and a limit L at a. Notice
that as we choose smaller values of ε (the distance between the function and the limit), we can always find a δ
small enough so that if we have chosen an x value within δ of a, then the value of f(x) is within ε of the limit
L.

Fig. 3.33 These graphs show possible values of δ, given successively smaller choices of ε
Activity 2
Visit the following applet to experiment with finding values of δ for selected values of ε:
 The epsilon-delta definition of limit (http://www.openstaxcollege.org/l/20_epsilondelt)

Example 3.31 show how you can use this definition to prove a statement about the limit of a specific function
at a specified value.
Example 3.31
Proving a Statement about the Limit of a Specific Function

Solution
Let ε> 0.
The first part of the definition begins “For every ε > 0.” This means we must prove that whatever
follows is true no matter what positive value of ε is chosen. By stating “Let ε > 0,” we signal our intent
to do so.

The definition continues with “there exists a δ > 0.” The phrase “there exists” in a mathematical
statement is always a signal for a scavenger hunt. In other words, we must go and find δ. So, where
exactly did δ = ε/2 come from? There are two basic approaches to tracking down δ. One method is
purely algebraic and the other is geometric.
We begin by tackling the problem from an algebraic point of view. Since ultimately we want |(2x + 1)
− 3| < ε, we begin by manipulating this expression: |(2x + 1) − 3| < ε is equivalent to |2x − 2| < ε,
which in turn is equivalent to |2||x − 1| < ε. Last, this is equivalent to |x − 1| < ε/2. Thus, it would seem
that δ = ε/2 is appropriate.
We may also find δ through geometric methods. Figure 3.34 demonstrates how this is done.
Fig. 3.34 This graph shows how we find δ geometrically

Analysis
In this part of the proof, we started with |(2x + 1) − 3| and used our assumption 0 < |x − 1| < δ in a key part
of the chain of inequalities to get |(2x + 1) − 3| to be less than ε. We could just as easily have
manipulated the assumed inequality 0 < |x − 1| < δ to arrive at |(2x + 1) − 3| < ε as follows:
The following Problem-Solving Strategy summarizes the type of proof we worked out in Example 3.31.
Example 3.32
Proving a Statement about a Limit

Solution

Activity 3
3.27 Complete the proof that by filling in the blanks.
In Example 3.31 and Example 3.32, the proofs were fairly straightforward, since the functions with which we
were working were linear. In Example 3.33, we see how to modify the proof to accommodate a nonlinear
function.

Example 3.33
Proving a Statement about the Limit of a Specific Function (Geometric Approach)

Prove that

Solution
1. Let ε > 0. The first part of the definition begins “For every ε > 0,” so we must prove that
whatever follows is true no matter what positive value of ε is chosen. By stating “Let ε >
0,” we signal our intent to do so.
2. Without loss of generality, assume ε ≤ 4. Two questions present themselves: Why do we
want ε ≤ 4 and why is it okay to make this assumption? In answer to the first question:
Later on, in the process of solving for δ, we will discover that δ involves the quantity 4
− ε. Consequently, we need ε ≤ 4. In answer to the second question: If we can find δ > 0
that “works” for ε ≤ 4, then it will “work” for any ε > 4 as well. Keep in mind that, although
it is always okay to put an upper bound on ε, it is never okay to put a lower bound (other
than zero) on ε.
3. Choose . Figure 3.35 shows how we made this choice of δ.

Fig. 3.35 This graph shows how we find δ geometrically for a given ε for the proof in Example 3.33

4. We must show: If 0 < |x − 2| < δ, then |x 2 − 4| < ε, so we must begin by assuming


0 < |x − 2| < δ.
We don’t really need 0 < |x − 2| (in other words, x ≠ 2) for this proof. Since 0 < |x − 2| < δ ⇒ |x − 2| < δ,
it is okay to drop 0 < |x − 2|.
|x − 2| < δ.
Hence,
−δ < x − 2 < δ.

Recall that and consequently


We also use here. We might ask at this point: Why did we
substitute for δ on the left-hand side of the inequality and on the right-hand
side of the inequality? If we look at Fig. 3.35, we see that corresponds to the distance
on the left of 2 on the x-axis and corresponds to the distance on the right. Thus,

The geometric approach to proving that the limit of a function takes on a specific value works quite well for some
functions. Also, the insight into the formal definition of the limit that this method provides is invaluable.
However, we may also approach limit proofs from a purely algebraic point of view. In many cases, an algebraic
approach may not only provide us with additional insight into the definition, it may prove to be simpler as well.
Furthermore, an algebraic approach is the primary tool used in proofs of statements about limits. For Example
3.34, we take on a purely algebraic approach.
Example 3.34
Proving a Statement about the Limit of a Specific Function (Algebraic Approach)

Solution
Let’s use our outline from the Problem-Solving Strategy:
1. Let ε > 0.
2. Choose δ = min{1, ε/5}. This choice of δ may appear odd at first glance, but it was obtained
by taking a look at our ultimate desired inequality: This inequality is

equivalent to At this point, the temptation simply to choose is


very strong. Unfortunately, our choice of δ must depend on ε only and no other variable. If we
can replace |x − 3| by a numerical value, our problem can be resolved. This is the place where
assuming δ ≤ 1 comes into play. The choice of δ ≤ 1 here is arbitrary. We could have just as
easily used any other positive number. In some proofs, greater care in this choice may be
necessary. Now, since δ ≤ 1 and |x + 1| < δ ≤ 1, we are able to show that |x − 3| < 5. Consequently,
|x + 1|· |x − 3| < |x + 1|· 5. At this point we realize that we also need δ ≤ ε/5. Thus, we choose
δ = min{1, ε/5}.
3. Assume 0 < |x + 1| < δ. Thus,
You will find that, in general, the more complex a function, the more likely it is that the algebraic approach is
the easiest to apply. The algebraic approach is also more useful in proving statements about limits.

Proving Limit Laws


We now demonstrate how to use the epsilon-delta definition of a limit to construct a rigorous proof of one of the
limit laws. The triangle inequality is used at a key point of the proof, so we first review this key property of
absolute value.

Definition
The triangle inequality states that if a and b are any real numbers, then |a + b| ≤ |a| + |b|.

Proof

We now explore what it means for a limit not to exist. The limit does not exist if there is no real number L

for which . Thus, for all real numbers L, . To understand what this means, we look
at each part of the definition of together with its opposite. A translation of the definition is
given in Table 3.8.
Table 3.8 Translation of the Definition of and its Opposite

Finally, we may state what it means for a limit not to exist. The limit does not exist if for every real
number L, there exists a real number ε > 0 so that for all δ > 0, there is an x satisfying 0 < |x − a| < δ, so that |f(x) − L|
≥ ε.

One-Sided and Infinite Limits


Just as we first gained an intuitive understanding of limits and then moved on to a more rigorous definition of a
limit, we now revisit one-sided limits. To do this, we modify the epsilon-delta definition of a limit to give formal
epsilon-delta definitions for limits from the right and left at a point. These definitions only require slight
modifications from the definition of the limit. In the definition of the limit from the right, the inequality
0 < x − a < δ replaces 0 < |x − a| < δ, which ensures that we only consider values of x that are greater than (to the
right of) a. Similarly, in the definition of the limit from the left, the inequality −δ < x − a < 0 replaces 0 < |x − a| < δ,
which ensures that we only consider values of x that are less than (to the left of) a.

Definition
Limit from the Right: Let f(x) be defined over an open interval of the form (a, b) where a < b. Then,

Example 3.35
Proving a Statement about a Limit From the Right

Prove that .

Solution
Let ε > 0.
Choose δ = ε2. Since we ultimately want we manipulate this inequality to get or,
equivalently, 0 < x − 4 < ε2, making δ = ε a clear choice. We may also determine δ geometrically, as
2

shown in Fig. 3.36.


Fig. 3.36 This graph shows how we find δ for the proof in Example 3.35

We conclude the process of converting our intuitive ideas of various types of limits to rigorous formal definitions
by pursuing a formal definition of infinite limits. To have , we want the values of the function f(x) to
get larger and larger as x approaches a. Instead of the requirement that |f(x) − L| < ε for arbitrarily small ε when
0 < |x − a| < δ for small enough δ, we want f(x) > M for arbitrarily large positive M when 0 < |x − a| < δ for small
enough δ. Figure 3.37 illustrates this idea by showing the value of δ for successively larger values of M.

Fig. 3.37 These graphs plot values of δ for M to show that

Definition
Let f(x) be defined for all x ≠ a in an open interval containing a. Then, we have an infinite limit

if for every M > 0, there exists δ > 0 such that if 0 < |x – a| < δ, then f(x) > M.
Let f(x) be defined for all x ≠ a in an open interval containing a. Then, we have a negative infinite limit

if for every M > 0, there exists δ > 0 such that if 0 < |x – a| < δ, then f(x) < −M.
Summary
 Differential calculus arose from trying to solve the problem of determining the slope of a line
tangent to a curve at a point. The slope of the tangent line indicates the rate of change of the
function, also called the derivative. Calculating a derivative requires finding a limit.
 Integral calculus arose from trying to solve the problem of finding the area of a region between
the graph of a function and the x-axis. We can approximate the area by dividing it into thin
rectangles and summing the areas of these rectangles. This summation leads to the value of a
function called the integral. The integral is also calculated by finding a limit and, in fact, is related
to the derivative of a function.
 Multivariable calculus enables us to solve problems in three-dimensional space, including
determining motion in space and finding volumes of solids.
 A table of values or graph may be used to estimate a limit.
 If the limit of a function at a point does not exist, it is still possible that the limits from the left and
right at that point may exist.
 If the limits of a function from the left and right exist and are equal, then the limit of the function
is that common value.
 For a function to be continuous at a point, it must be defined at that point, its limit must exist at
the point, and the value of the function at that point must equal the value of the limit at that point.
 Discontinuities may be classified as removable, jump, or infinite.
 A function is continuous over an open interval if it is continuous at every point in the interval. It
is continuous over a closed interval if it is continuous at every point in its interior and is
continuous at its endpoints.

Keywords
 Average Velocity the change in an object’s position divided by the length of a time period; the
average velocity of an object over a time interval , with a

position given by s(t), that is


 Constant Multiple Law for Limits the limit law
 Differential Calculus the field of calculus concerned with the study of derivatives and their
applications
 Discontinuity at a Point A function is discontinuous at a point or has a discontinuity at a point
if it is not continuous at the point
 Infinite Limit A function has an infinite limit at a point a if it either increases or decreases
without bound as it approaches a

Self-Assessment Questions
1. Estimate the slope of the tangent line (rate of change) to f(x) = x2 at x = 1 by finding slopes of
secant lines through (1, 1) and the point on the graph of f(x) = x2.
2. Evaluate each of the following limits. Identify any vertical asymptotes of the function
3. Evaluate for f (x) shown here:

4. Evaluate

5. Evaluate

6. Use the squeeze theorem to evaluate

7. Using the definition, determine whether the function is continuous at x = 1. If


the function is not continuous at 1, indicate the condition for continuity at a point that fails to
hold.

8. Evaluate
9. Find δ corresponding to ε > 0 for a proof that .
10. A ball is thrown into the air and the vertical position is given by x(t) = −4.9t2 + 25t + 5. Use
the Intermediate Value Theorem to show that the ball must land on the ground sometime
between 5 sec and 6 sec after the throw.
Unit 4
Derivatives
Structure
4.1 Introduction
4.2 Defining the Derivative
4.3 The Derivative as a Function
4.4 Differentiation Rules
4.5 Derivatives as Rates of Change
4.6 Derivatives of Trigonometric Functions
4.7 The Chain Rule
4.8 Implicit Differentiation
4.9 Derivatives of Exponential and Logarithmic Functions
Summary
Keywords
Self-Assessment Questions

The text is adapted by Symbiosis Centre for Distance Learning under a Creative Commons
Attribution-NonCommercial-ShareAlike 4.0 International (CC BY-NC-SA 4.0) as requested by the
work’s creator or licensees. This license is available at https://creativecommons.org/licenses/by-nc-
sa/4.0/
Objectives
After going through this unit, you will be able to:
 Define the derivative function of a given function
 Explain the meaning of a higher-order derivative
 State the constant, constant multiple, and power rules
 Apply the sum and difference rules to combine derivatives
 Use the product rule for finding the derivative of a product of functions
 Determine a new value of a quantity from the old value and the amount of change
 Find the derivative of exponential and logarithmic functions
 Use logarithmic differentiation to determine the derivative of a function

4.1 INTRODUCTION
The Hennessey Venom GT is one of the fastest cars in the world. In 2014, it reached a record-setting speed of
270.49 mph. It can go from 0 to 200 mph in 14.51 seconds. The techniques in this chapter can be used to
calculate the acceleration the Venom achieves in this feat.
Calculating velocity and changes in velocity are important uses of calculus, but it is far more widespread than
that. Calculus is important in all branches of mathematics, science, and engineering, and it is critical to analysis
in business and health as well. In this chapter, we explore one of the main tools of calculus, the derivative, and
show convenient ways to calculate derivatives. We apply these rules to a variety of functions in this chapter so
that we can then explore applications of these techniques.

4.2 DEFINING THE DERIVATIVE


Tangent Lines
We begin our study of calculus by revisiting the notion of secant lines and tangent lines. Recall that we used
the slope of a secant line to a function at a point (a, f(a)) to estimate the rate of change, or the rate at which one
variable changes in relation to another variable. We can obtain the slope of the secant by choosing a value of x
near a and drawing a line through the points (a, f(a)) and (x, f(x)), as shown in Fig. 4.1. The slope of this line is
given by an equation in the form of a difference quotient:

We can also calculate the slope of a secant line to a function at a value a by using this equation and replacing
x with a + h, where h is a value close to 0. We can then calculate the slope of the line through the points (a,
f(a)) and (a + h, f(a + h)). In this case, we find the secant line has a slope given by the following difference
quotient with increment h:
Definition
Let f be a function defined on an interval I containing a. If x ≠ a is in I, then

(4.1)
is a difference quotient.
Also, if h ≠ 0 is chosen so that a + h is in I, then

(4.2)
is a difference quotient with increment h.

These two expressions for calculating the slope of a secant line are illustrated in Fig. 4.1. We will see that
each of these two methods for finding the slope of a secant line is of value. Depending on the setting, we can
choose one or the other. The primary consideration in our choice usually depends on ease of calculation.

Fig. 4.1 We can calculate the slope of a secant line in either of two ways

In Fig. 4.2(a) we see that, as the values of x approach a, the slopes of the secant lines provide better
estimates of the rate of change of the function at a. Furthermore, the secant lines themselves approach the
tangent line to the function at a, which represents the limit of the secant lines. Similarly, Fig. 4.2(b) shows
that as the values of h get closer to 0, the secant lines also approach the tangent line. The slope of the
tangent line at a is the rate of change of the function at a, as shown in Fig. 4.2(c).
Fig. 4.2 The secant lines approach the tangent line (shown in green) as the second point approaches the
first

In Fig. 4.3 we show the graph of and its tangent line at (1, 1) in a series of tighter intervals
about x = 1.
As the intervals become narrower, the graph of the function and its tangent line appear to coincide, making
the values on the tangent line a good approximation to the values of the function for choices of x close to
1. In fact, the graph of f(x) itself appears to be locally linear in the immediate vicinity of x = 1.

Fig. 4.3 For values of x close to 1, the graph of f (x) = x and its tangent line appear to coincide
Formally we may define the tangent line to the graph of a function as follows.

Definition
Let f (x) be a function defined in an open interval containing a. The tangent line to f (x) at a is the line
passing through the point (a, f(a)) having slope

(4.3)
provided this limit exists.
Equivalently, we may define the tangent line to f (x) at a to be the line passing through the point (a, f(a))
having slope

(4.4)
provided this limit exists.

Just as we have used two different expressions to define the slope of a secant line, we use two different
forms to define the slope of the tangent line. In this text we use both forms of the definition. As before, the
choice of definition will depend on the setting. Now that we have formally defined a tangent line to a
function at a point, we can use this definition to find equations of tangent lines.

Example 4.1

Finding a Tangent Line


Find the equation of the line tangent to the graph of f (x) = x2 at x = 3.

Solution

First find the slope of the tangent line. In this example, use Equation 4.3.

Using the point-slope equation of the line with the slope m = 6 and the point (3, 9), we obtain the
line y – 9 = 6(x − 3). Simplifying, we have y = 6x − 9. The graph of f (x) = x 2 and its tangent line
at 3 are shown in Fig. 4.4.
Fig. 4.4 The tangent line to f (x) at x = 3.

Example 4.2

The Slope of a Tangent Line Revisited

Use Equation 4.4 to find the slope of the line tangent to the graph of f (x) = x2 at x = 3.

Solution

The steps are very similar to Example 4.1. See Equation 4.4 for the definition.

Example 4.3

Finding the Equation of a Tangent Line

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

Solution

We can use Equation 4.3, but as we have seen, the results are the same if we use Equation 4.4.
Fig. 4.5.

Fig. 4.5 The line is tangent to f (x) at x = 2

The Derivative of a Function at a point


The type of limit we compute in order to find the slope of the line tangent to a function at a point occurs in
many applications across many disciplines. These applications include velocity and acceleration in physics,
marginal profit functions in business, and growth rates in biology. This limit occurs so frequently that we
give this value a special name: the derivative. The process of finding a derivative is called differentiation.

Definition
Let f(x) be a function defined in an open interval containing a. The derivative of the function f(x) at a,
denoted by f ′(a), is defined by

(4.5)
provided this limit exists.
Alternatively, we may also define the derivative of f(x) at a as
(4.6)

Example 4.4

Estimating a Derivative

For f (x) = x2, use a table to estimate f ′(3) using Equation 4.5.

Solution

Create a table using values of x just below 3 and just above 3.

After examining the table, we see that a good estimate is f ′ (3) = 6.

Example 4.5

Finding a Derivative

For f (x) = 3x2 − 4x + 1, find f ′(2) by using Equation 4.5.

Solution

Substitute the given function and value directly into the equation.
Example 4.6

Revisiting the Derivative

For f (x) = 3x2 − 4x + 1, find f ′(2) by using Equation 4.6.

Solution

Using this equation, we can substitute two values of the function into the equation, and we should
get the same value as in Example 4.5.

The results are the same whether we use Equation 4.5 or Equation 4.6.

Velocities and Rates of Change


Now that we can evaluate a derivative, we can use it in velocity applications. Recall that if s(t) is the
position of an object moving along a coordinate axis, the average velocity of the object over a time interval
[a, t] if t > a or [t, a] if t < a is given by the difference quotient

(4.7)
As the values of t approach a, the values of vave approach the value we call the instantaneous velocity at
a. That is, instantaneous velocity at a, denoted v(a), is given by

(4.8)
To better understand the relationship between average velocity and instantaneous velocity, see Fig. 4.6. In
this figure, the slope of the tangent line (shown in red) is the instantaneous velocity of the object at time t
= a whose position at time t is given by the function s(t). The slope of the secant line (shown in green) is
the average velocity of the object over the time interval [a, t].

Fig. 4.6 The slope of the secant line is the average velocity over the interval [a, t]. The slope of the
tangent line is the instantaneous velocity

We can use Equation 4.5 to calculate the instantaneous velocity, or we can estimate the velocity of a
moving object by using a table of values. We can then confirm the estimate by using Equation 4.7.

Definition
The instantaneous rate of change of a function f (x) at a value a is its derivative f ′(a).

Example 4.7

Rate of Change of Temperature

A homeowner sets the thermostat so that the temperature in the house begins to drop from 70°F at
9 p.m., reaches a low of 60° during the night, and rises back to 70° by 7 a.m. the next morning.
Suppose that the temperature in the house is given by T(t) = 0.4t2 − 4t + 70 for 0 ≤ t ≤ 10, where t is
the number of hours past 9 p.m. Find the instantaneous rate of change of the temperature at
midnight.

Solution

Since midnight is 3 hours past 9 p.m., we want to compute T′(3). Refer to Equation 4.5.
4.3 THE DERIVATIVE AS A FUNCTION
As we have seen, the derivative of a function at a given point gives us the rate of change or slope of the
tangent line to the function at that point. If we differentiate a position function at a given time, we obtain
the velocity at that time. It seems reasonable to conclude that knowing the derivative of the function at
every point would produce valuable information about the behavior of the function. However, the process
of finding the derivative at even a handful of values using the techniques of the preceding section would
quickly become quite tedious. In this section we define the derivative function and learn a process for
finding it.
Derivative Functions
The derivative function gives the derivative of a function at each point in the domain of the original
function for which the derivative is defined. We can formally define a derivative function as follows.

Definition
Let f be a function. The derivative function, denoted by f ′, is the function whose domain consists of those values
of x such that the following limit exists:

(4.9)

A function f (x) is said to be differentiable at a if f ′(a) exists. More generally, a function is said to be
differentiable on S if it is differentiable at every point in an open set S, and a differentiable function is
one in which f ′(x) exists on its domain.
In the next few examples we use Equation 4.9 to find the derivative of a function.
Example 4.8

Finding the Derivative of a Square-Root Function


Solution

Start directly with the definition of the derivative function. Use Equation 4.1.

Example 4.9

Finding the Derivative of a Quadratic Function

Find the derivative of the function f (x) = x2 − 2x.

Solution

We use a variety of different notations to express the derivative of a function. In Example 4.12 we showed
that if f (x) = x2 − 2x, then f ′ (x) = 2x − 2. If we had expressed this function in the form y = x2 − 2x, we could

have expressed the derivative as y′ = 2x − 2 or . We could have conveyed the same information by

writing Thus, for the function y = f(x), each of the following notations represents the
derivative of f(x):
In place of f ′(a) we may also use Use of the notation (called Leibniz notation) is quite common
in engineering and physics. To understand this notation better, recall that the derivative of a function at a point
is the limit of the slopes of secant lines as the secant lines approach the tangent line. The slopes of these secant

lines are often expressed in the form where Δy is the difference in the y values corresponding to the
difference in the x values, which are expressed as Δx (Fig. 4.7). Thus the derivative, which can be thought of
as the instantaneous rate of change of y with respect to x, is expressed as

Fig. 4.7 The derivative is expressed as

4.4 DIFFERENTIATION RULES


Finding derivatives of functions by using the definition of the derivative can be a lengthy and, for certain

functions, a rather challenging process. For example, previously we found that by using a
process that involved multiplying an expression by a conjugate prior to evaluating a limit. The process that we
could use to evaluate using the definition, while similar, is more complicated. In this section, we
develop rules for finding derivatives that allow us to bypass this process. We begin with the basics.
The Basic Rules
The functions f(x) = c and g(x) = xn where n is a positive integer are the building blocks from which all
polynomials and rational functions are constructed. To find derivatives of polynomials and rational functions
efficiently without resorting to the limit definition of the derivative, we must first develop formulas for
differentiating these basic functions.
The Constant Rule
We first apply the limit definition of the derivative to find the derivative of the constant function, f (x) = c. For
this function, both f (x) = c and f (x + h) = c, so we obtain the following result:

The rule for differentiating constant functions is called the constant rule. It states that the derivative of a
constant function is zero; that is, since a constant function is a horizontal line, the slope, or the rate of change,
of a constant function is 0. We restate this rule in the following theorem.

Theorem 4.1: The Constant Rule

Example 4.10

Applying the Constant Rule

Find the derivative of f (x) = 8.

Solution

This is just a one-step application of the rule:

f ′(8) = 0.
The Power Rule
We have shown that

At this point, you might see a pattern beginning to develop for derivatives of the form . We continue our

examination of derivative formulas by differentiating power functions of the form f (x) = xn where n is a positive
integer. We develop formulas for derivatives of this type of function in stages, beginning with positive integer
powers. Before stating and proving the general rule for derivatives of functions of this form, we take a look at

a specific case, . As we go through this derivation, pay special attention to the portion of the expression
in boldface, as the technique used in this case is essentially the same as the technique used to prove the general
case.
Example 4.11
Differentiating x3

Solution

As we shall see, the procedure for finding the derivative of the general form f (x) = xn is very similar. Although
it is often unwise to draw general conclusions from specific examples, we note that when we differentiate
f(x) = x 3, the power on x becomes the coefficient of x2 in the derivative and the power on x in the derivative
decreases by 1. The following theorem states that the power rule holds for all positive integer powers of x. We
will eventually extend this result to negative integer powers. Later, we will see that this rule may also be
extended first to rational powers of x and then to arbitrary powers of x. Be aware, however, that this rule
does not apply to functions in which a constant is raised to a variable power, such as f(x) = 3x.
Theorem 4.2: The Power Rule

Proof
For f(x) = xn where n is a positive integer, we have

We see that

Example 4.12
Applying the Power Rule

Find the derivative of the function f (x) = x10 by applying the power rule.

Solution
Using the power rule with n = 10, we obtain
f ′(x) = 10x10 − 1 = 10x9.

The Sum, Difference, and Constant Multiple Rules


We find our next differentiation rules by looking at derivatives of sums, differences, and constant multiples of
functions. Just as when we work with functions, there are rules that make it easier to find derivatives of functions
that we add, subtract, or multiply by a constant. These rules are summarized in the following theorem.
Theorem 4.3: Sum, Difference, and Constant Multiple Rules

Let f (x) and g(x) be differentiable functions and k be a constant. Then each of the following equations holds.

Sum Rule. The derivative of the sum of a function f and a function g is the same as the sum of the derivative of f and the
derivative of g.

that is,

Difference Rule. The derivative of the difference of a function f and a function g is the same as the difference of the
derivative of f and the derivative of g:

Proof
We provide only the proof of the sum rule here. The rest follow in a similar manner.
For differentiable functions f (x) and g(x), we set j(x) = f (x) + g(x). Using the limit definition of the derivative we have

Example 4.13

Applying the Constant Multiple Rule

Find the derivative of g(x) = 3x2 and compare it to the derivative of f (x) = x2.
Solution
Since f (x) = x2 has derivative f ′ (x) = 2x, we see that the derivative of g(x) is 3 times the
derivative of f (x). This relationship is illustrated in Fig. 4.8.

Fig. 4.8 The derivative of g(x) is 3 times the derivative of f(x)

Example 4.14

Applying Basic Derivative Rules

Find the derivative of f (x) = 2x5 + 7.

Solution

We begin by applying the rule for differentiating the sum of two functions, followed by the rules
for differentiating constant multiples of functions and the rule for differentiating powers. To better
understand the sequence in which the differentiation rules are applied, we use Leibniz notation
throughout the solution:

Example 4.15

Finding the Equation of a Tangent Line


Find the equation of the line tangent to the graph of f(x) = x2 − 4x + 6 at x = 1.

Solution

The Product Rule


Now that we have examined the basic rules, we can begin looking at some of the more advanced rules. The first one
examines the derivative of the product of two functions. Although it might be tempting to assume that the derivative
of the product is the product of the derivatives, similar to the sum and difference rules, the product rule does not
follow this pattern. To see why we cannot use this pattern, consider the function f (x) = x2, whose derivative is f ′ (x) =

2x and not

Theorem 4.4: Product Rule


Let f(x) and g(x) be differentiable functions. Then

This means that the derivative of a product of two functions is the derivative of the first function times the
second function plus the derivative of the second function times the first function.

Proof

→0
Example 4.16

Applying the Product Rule to Functions at a Point

Solution

Example 4.17

Applying the Product Rule to Binomials

Solution
The Quotient Rule
Having developed and practiced the product rule, we now consider differentiating quotients of functions. As we see
in the following theorem, the derivative of the quotient is not the quotient of the derivatives; rather, it is the derivative
of the function in the numerator times the function in the denominator minus the derivative of the function in the
denominator times the function in the numerator, all divided by the square of the function in the denominator. In
order to better grasp why we cannot simply take the quotient of the derivatives, keep in mind that

Theorem 4.5: The Quotient Rule

The proof of the quotient rule is very similar to the proof of the product rule, so it is omitted here. Instead, we
apply this new rule for finding derivatives in the next example.
Example 4.18

Applying the Quotient Rule

Solution

It is now possible to use the quotient rule to extend the power rule to find derivatives of functions of the form xk
where k is a negative integer.
Theorem 4.6: Extended Power Rule
If k is a negative integer, then

Proof

Example 4.19

Using the Extended Power Rule

Solution

By applying the extended power rule with k = −4, we obtain

Example 4.20

Using the Extended Power Rule and the Constant Multiple Rule

2
Solution

Combining Differentiation Rules


As we have seen throughout the examples in this section, it seldom happens that we are called on to apply just one
differentiation rule to find the derivative of a given function. At this point, by combining the differentiation rules, we
may find the derivatives of any polynomial or rational function. Later on we will encounter more complex
combinations of differentiation rules. A good rule of thumb to use when applying several rules is to apply the rules
in reverse of the order in which we would evaluate the function.

4.5 DERIVATIVES AS RATES OF CHANGE


In this section, we look at some applications of the derivative by focusing on the interpretation of the derivative as the
rate of change of a function. These applications include acceleration and velocity in physics, population growth
rates in biology, and marginal functions in economics.
Amount of Change Formula

(4.10)
We can use this formula if we know only f (a) and f ′(a) and wish to estimate the value of f (a + h). For example,
we may use the current population of a city and the rate at which it is growing to estimate its population in the
near future. As we can see in Fig. 4.9, we are approximating f (a + h) by the y coordinate at a + h on the line
tangent to f (x) at x = a. Observe that the accuracy of this estimate depends on the value of h as well as the value
of f ′(a).

Fig. 4.9 The new value of a changed quantity equals the original value plus the rate of change times the interval of

change:
Motion along a Line
Another use for the derivative is to analyze motion along a line. We have described velocity as the rate of change of
position. If we take the derivative of the velocity, we can find the acceleration, or the rate of change of velocity. It is
also important to introduce the idea of speed, which is the magnitude of velocity. Thus, we can state the following
mathematical definitions.

Definition

Changes in Cost and Revenue


In addition to analyzing motion along a line and population growth, derivatives are useful in analyzing changes in
cost, revenue, and profit. The concept of a marginal function is common in the fields of business and economics and
implies the use of derivatives. The marginal cost is the derivative of the cost function. The marginal revenue is the
derivative of the revenue function. The marginal profit is the derivative of the profit function, which is based on the
cost function and the revenue function.
Definition

Example 4.21

Applying Marginal Revenue

Assume that the number of barbeque dinners that can be sold, x, can be related to the price charged,
p, by the equation p(x) = 9 − 0.03x, 0 ≤ x ≤ 300.

In this case, the revenue in dollars obtained by selling x barbeque dinners is given by

R(x) = xp(x) = x(9 − 0.03x) = −0.03x2 + 9x for 0 ≤ x ≤ 300.

Use the marginal revenue function to estimate the revenue obtained from selling the 101st
barbeque dinner. Compare this to the actual revenue obtained from the sale of this dinner.

Solution

First, find the marginal revenue function: MR(x) = R′ (x) = −0.06x + 9.

Next, use R′ (100) to approximate R(101) − R(100), the revenue obtained from the sale of the 101st
dinner. Since R′ (100) = 3, the revenue obtained from the sale of the 101st dinner is approximately
$3.

The actual revenue obtained from the sale of the 101st dinner is

R(101) − R(100) = 602.97 − 600 = 2.97, or $2.97.

The marginal revenue is a fairly good estimate in this case and has the advantage of being easy to
compute.

4.6 DERIVATIVES OF TRIGONOMETRIC FUNCTIONS


One of the most important types of motion in physics is simple harmonic motion, which is associated with such
systems as an object with mass oscillating on a spring. Simple harmonic motion can be described by using either sine
or cosine functions. In this section we expand our knowledge of derivative formulas to include derivatives of these
and other trigonometric functions. We begin with the derivatives of the sine and cosine functions and then use them to
obtain formulas for the derivatives of the remaining four trigonometric functions. Being able to calculate the
derivatives of the sine and cosine functions will enable us to find the velocity and acceleration of simple harmonic
motion.
Derivatives of the Sine and Cosine Functions

By setting and using a graphing utility, we can get a graph of an approximation to the
derivative of sin x (Fig. 4.10).

Fig. 4.10 The graph of the function D(x) looks a lot like a cosine curve
Theorem 4.7: The Derivatives of sin x and cos x
The derivative of the sine function is the cosine and the derivative of the cosine function is the negative sine.

(4.11)

(4.12)

Proof

Because the proofs for use similar techniques, we provide only the

proof for Before beginning, recall two important trigonometric limits we learned in
Limits:

The graphs of are shown in Fig. 4.11.

Fig. 4.11 These graphs show two important limits needed to establish the derivative formulas for the sine and
cosine functions
Figure 4.12 shows the relationship between the graph of f (x) = sin x and its derivative f ′ (x) = cos x. Notice
that at the points where f (x) = sin x has a horizontal tangent, its derivative f ′ (x) = cos x takes on the value
zero. We also see that where f (x) = sin x is increasing, f ′ (x) = cos x > 0 and where f (x) = sin x is decreasing,
f ′ (x) = cos x < 0.

Fig. 4.12 Where f (x) has a maximum or a minimum, f ′(x) = 0 that is, f ′(x) = 0 where f (x) has a
horizontal tangent. These points are noted with dots on the graphs

Example 4.22

Differentiating a Function Containing sin x

Find the derivative of f (x) = 5x3 sin x.


Solution

Example 4.23

Finding the Derivative of a Function Containing cos x

Solution

By applying the quotient rule, we have

Simplifying, we obtain

Derivatives of Other Trigonometric Functions

Since the remaining four trigonometric functions may be expressed as quotients involving sine, cosine, or
both, we can use the quotient rule to find formulas for their derivatives.

Theorem 4.8: Derivatives of tanx, cot x, sec x, and csc x

The derivatives of the remaining trigonometric functions are as follows:

(4.13)

(4.14)

(4.15)
(4.16)

Example 4.24

Finding the Derivative of Trigonometric Functions

Find the derivative of f (x) = csc x + x tan x.

Solution

To find this derivative, we must use both the sum rule and the product rule. Using the sum rule,
we find

4.7 THE CHAIN RULE

Rule: The Chain Rule

(4.17)
Activity 1
Watch an animation (http://www.openstaxcollege.org/l/20_chainrule2) of the chain rule.

The Chain and Power Rules Combined

Rule: Power Rule for Composition of Functions

For all values of x for which the derivative is defined, if

Then

(4.18)

4.8 IMPLICIT DIFFERENTIATION


We have already studied how to find equations of tangent lines to functions and the rate of change of a function
at a specific point. In all these cases we had the explicit equation for the function and differentiated these
functions explicitly. Suppose instead that we want to determine the equation of a tangent line to an arbitrary
curve or the rate of change of an arbitrary curve at a point. In this section, we solve these problems by finding
the derivatives of functions that define y implicitly in terms of x.
Implicit Differentiation
In most discussions of math, if the dependent variable y is a function of the independent variable x, we express
y in terms of x. If this is the case, we say that y is an explicit function of x. For example, when we write the
equation y = x2 + 1, we are defining y explicitly in terms of x. On the other hand, if the relationship between the
function y and the variable x is expressed by an equation where y is not expressed entirely in terms of x, we say
that the equation defines y implicitly in terms of x. For example, the equation y − x 2 = 1 defines the function y =
x 2 + 1 implicitly.
Implicit differentiation allows us to find slopes of tangents to curves that are clearly not functions (they fail the
vertical line test). We are using the idea that portions of y are functions that satisfy the given equation, but that
y is not actually a function of x.
In general, an equation defines a function implicitly if the function satisfies that equation. An equation may

define many different functions implicitly. For example, the functions and

which are illustrated in Fig. 4.13, are just three of the many functions defined
implicitly by the equation x2 + y2 = 25.

Fig. 4.13 The equation x2 + y2 = 25 defines many functions implicitly


Example 4.25

Using Implicit Differentiation

Solution
Analysis

4.9 DERIVATIVES OF EXPONENTIAL AND LOGARITHMIC FUNCTIONS


So far, we have learned how to differentiate a variety of functions, including trigonometric, inverse, and implicit
functions. In this section, we explore derivatives of exponential and logarithmic functions. As we know,
exponential functions play an important role in modeling population growth and the decay of radioactive
materials. Logarithmic functions can help rescale large quantities and are particularly helpful for rewriting
complicated expressions.
Derivative of the Exponential Function
Just as when we found the derivatives of other functions, we can find the derivatives of exponential and
logarithmic functions using formulas. As we develop these formulas, we need to make certain basic
assumptions. The proofs that these assumptions hold are beyond the scope of this course.
Table 4.1 Approximating a Value of 4π

We also assume that for B(x) = bx, b > 0, the value B′ (0) of the derivative exists. In this section, we show that
by making this one additional assumption, it is possible to prove that the function B(x) is differentiable
everywhere.
We make one final assumption: that there is a unique value of b > 0 for which B′ (0) = 1. We define e to be this

unique value. Figure 4.14 provides graphs of the functions y = 2 x, y = 3x, y = 2.7 x, and y = 2.8 x. A visual
estimate of the slopes of the tangent lines to these functions at 0 provides evidence that the value of e lies

somewhere between 2.7 and 2.8. The function E(x) = e x is called the natural exponential function. Its inverse,
L(x) = loge x = lnx is called the natural logarithmic function.

Fig. 4.14 The graph of E(x) = e x is between y = 2 x and y = 3x.


Table 4.2 Estimating a Value of e

The evidence from the table suggests that 2.7182 < e < 2.7183.

The graph of E(x) = ex together with the line y = x + 1 are shown in Fig. 4.15. This line is tangent to the graph

of E(x) = ex at x = 0.

Fig. 4.15 The tangent line to E(x) = ex at x = 0 has slope

Now that we have laid out our basic assumptions, we begin our investigation by exploring the

derivative of B(x) = bx, b > 0. Recall that we have assumed that B′ (0) exists. By applying the limit definition
to the derivative we conclude that

(4.19)
Turning to B′ (x), we obtain the following.

We see that on the basis of the assumption that B(x) = b x is differentiable at 0, B(x) is not only differentiable
everywhere, but its derivative is

(4.20)

For E(x) = ex, E′ (0) = 1. Thus, we have E′ (x) = ex. (The value of B′ (0) for an arbitrary function of the form

B(x) = bx, b > 0, will be derived later.)

Theorem 4.9: Derivative of the Natural Exponential Function

Example 4.26
Example 4.27

Example 4.28

If A(t) = 1000e0.3t describes the mosquito population after t days, as in the preceding example, what is the rate
of change of A(t) after 4 days?
Derivative of the Logarithmic Function
Now that we have the derivative of the natural exponential function, we can use implicit differentiation to find
the derivative of its inverse, the natural logarithmic function.

Theorem 4.10: The Derivative of the Natural Logarithmic Function


If x > 0 and y = ln x, then

(4.21)
More generally, let g(x) be a differentiable function. For all values of x for which g′ (x) > 0, the derivative
of h(x) = ln (g(x))is given by

(4.22)

Proof

Finally, we substitute x = ey to obtain

The graph of y = ln x and its derivative are shown in Fig. 4.36.

Fig. 4.16 The function y = lnx is increasing on (0, +∞). Its derivative is greater than zero on (0, +∞)
Example 4.29

Taking a Derivative of a Natural Logarithm

Solution

Use Equation 4.22 directly.

Now that we can differentiate the natural logarithmic function, we can use this result to find the derivatives of

y = logbx and y = bx for b > 0, b ≠ 1.

Theorem 4.11: Derivatives of General Exponential and Logarithmic Functions


Let b > 0, b ≠ 1, and let g(x) be a differentiable function.
i. If, y = logb x, then

(4.23)
More generally, if h(x) = logb(g(x)), then for all values of x for which g(x)>0,

(4.24)
ii. If y = bx, then

(4.25)
More generally, if h(x) = bg(x), then

(4.26)

Proof
The derivative in Equation 4.24 now follows from the chain rule.

The more general derivative (Equation 4.26) follows from the chain rule.
Example 4.30

Applying Derivative Formulas

Solution

Use the quotient rule and Derivatives of General Exponential and Logarithmic Functions.

Example 4.31

Finding the Slope of a Tangent Line

Find the slope of the line tangent to the graph of y = log2(3x + 1) at x = 1.

Solution

To find the slope, we must evaluate Using Equation 4.24, we see that
Logarithmic Differentiation

Example 4.32
Summary

 The slope of the tangent line to a curve measures the instantaneous rate of change of a curve. We can
calculate it by finding the limit of the difference quotient or the difference quotient with increment h.

 The derivative of a function f (x) at a value a is found using either of the definitions for the slope of the
tangent line.

 Velocity is the rate of change of position. As such, the velocity v(t) at time t is the derivative of the
position s(t) at time t. Average velocity is given by

Instantaneous velocity is given by

 The derivative of a function f (x) is the function whose value at x is f ′(x).

 The graph of a derivative of a function f (x) is related to the graph of f (x). Where f (x) has a tangent
line with positive slope, f ′ (x) > 0. Where f (x) has a tangent line with negative slope, f ′ (x) < 0. Where
f (x) has a horizontal tangent line, f ′(x) = 0.

 If a function is differentiable at a point, then it is continuous at that point. A function is not


differentiable at a point if it is not continuous at the point, if it has a vertical tangent line at the point,
or if the graph has a sharp corner or cusp.

 Higher-order derivatives are derivatives of derivatives, from the second derivative to the nth derivative.

 The derivative of a constant function is zero.

 The derivative of a power function is a function in which the power on x becomes the coefficient of
the term and the power on x in the derivative decreases by 1.

 The derivative of a constant c multiplied by a function f is the same as the constant multiplied by the
derivative.

 The derivative of the sum of a function f and a function g is the same as the sum of the derivative of f
and the derivative of g.

 The derivative of the difference of a function f and a function g is the same as the difference of the
derivative of f
and the derivative of g.
 The derivative of a product of two functions is the derivative of the first function times the second
function plus the derivative of the second function times the first function.

 The derivative of the quotient of two functions is the derivative of the first function times the second
function minus the derivative of the second function times the first function, all divided by the square
of the second function.

 We used the limit definition of the derivative to develop formulas that allow us to find derivatives
without resorting to the definition of the derivative. These formulas can be used singly or in
combination with each other.

 The rate of change of position is velocity, and the rate of change of velocity is acceleration. Speed is
the absolute value, or magnitude, of velocity.

 The population growth rate and the present population can be used to predict the size of a future
population.

 Marginal cost, marginal revenue, and marginal profit functions can be used to predict, respectively, the
cost of producing one more item, the revenue obtained by selling one more item, and the profit obtained
by producing and selling one more item.
Keywords
 Acceleration is the rate of change of the velocity, that is, the derivative of velocity.

 Amount of change the amount of a function f (x) over an interval .

 Average rate of change is a function f (x) over an interval .

 Chain rule the chain rule defines the derivative of a composite function as the derivative of the
outer function evaluated at the inner function times the derivative of the inner function.
 Constant multiple rule the derivative of a constant c multiplied by a function f is the same as the

constant multiplied by the derivative .

 Constant rule the derivative of a constant function is zero: , where c is a constant.

 Derivative the slope of the tangent line to a function at a point, calculated by taking the limit of the
difference quotient, is the derivative.

 Derivative function gives the derivative of a function at each point in the domain of the original
function for which the derivative is defined.
 Difference quotient of a function f (x) at a is given by

 Difference rule the derivative of the difference of a function f and a function g is the same as the

difference of the derivative of f and the derivative of g: .

 Differentiable at a a function for which f ′(a) exists is differentiable at a.

 Differentiable function a function for which f ′(x) exists is a differentiable function.

 Differentiable on S a function for which f ′(x) exists for each x in the open set S is differentiable on S.

 Differentiation the process of taking a derivative.

 Higher-order derivative a derivative of a derivative, from the second derivative to the nth derivative,
is called a higher- order derivative

Self-Assessment Questions

1. For f (x) = x2+ 3x + 2, find f ′ (1).

2. Find the derivative of functions given below.


i. g(x) = −3.
ii. f (x) = x7.
iii. f (x) = 2x3 − 6x2 + 3.

iv.
v. f (x) = sin x cos x.
vi. h(x) = xe2x.
3. Find the equation of the line tangent to the graph of f (x) = 3x2 − 11 at x = 2. Use the point-slope form.

4. Use the product rule to obtain the derivative of j(x) = 2x5(4x2 + x).

5. Find the derivative of using the extended power rule.

6. Find the slope for the line tangent to y = 3x at x = 2.


Unit 5
Integration
Structure
5.1 Introduction
5.2 Approximating Areas
5.3 The Definite Integral
5.4 Integration Formulas
5.5 Substitution
Summary
Keywords
Self-Assessment Questions

The text is adapted by Symbiosis Centre for Distance Learning under a Creative Commons
Attribution-NonCommercial-ShareAlike 4.0 International (CC BY-NC-SA 4.0) as requested by the
work’s creator or licensees. This license is available at https://creativecommons.org/licenses/by-nc-
sa/4.0/
Objectives
After going through this unit, you will be able to:
 Use sigma (summation) notation to calculate sums and powers of integers

 Use the sum of rectangular areas to approximate the area under a curve

 Explain the terms integrand, limits of integration, and variable of integration

 Explain when a function is integrable

 Describe the relationship between the definite integral and net area

 To study the use of substitution to evaluate indefinite integrals

 To know how substitution used to evaluate definite integrals


5.1 INTRODUCTION
Determining distance from velocity is just one of many applications of integration. In fact, integrals are used in
a wide variety of mechanical and physical applications. In this chapter, we first introduce the theory behind
integration and use integrals to calculate areas. From there, we develop the Fundamental Theorem of Calculus,
which relates differentiation and integration. We then study some basic integration techniques and briefly
examine some applications.
5.2 APPROXIMATING AREAS
Archimedes was fascinated with calculating the areas of various shapes—in other words, the amount of space
enclosed by the shape. He used a process that has come to be known as the method of exhaustion, which used
smaller and smaller shapes, the areas of which could be calculated exactly, to fill an irregular region and thereby
obtain closer and closer approximations to the total area. In this process, an area bounded by curves is filled
with rectangles, triangles, and shapes with exact area formulas. These areas are then summed to approximate
the area of the curved region.
In this section, we develop techniques to approximate the area between a curve, defined by a function f (x), and
the x-axis on a closed interval [a, b]. Like Archimedes, we first approximate the area under the curve using
shapes of known area (namely, rectangles). By using smaller and smaller rectangles, we get closer and closer
approximations to the area. Taking a limit allows us to calculate the exact area under the curve.
Let’s start by introducing some notation to make the calculations easier. We then consider the case when f (x)
is continuous and nonnegative.
Sigma (Summation) Notation
As mentioned, we will use shapes of known area to approximate the area of an irregular region bounded by
curves. This process often requires adding up long strings of numbers. To make it easier to write down these
lengthy sums, we look at some new notation here, called sigma notation (also known as summation notation).
The Greek capital letter Σ, sigma, is used to express long sums of values in a compact form. For example, if we
want to add all the integers from 1 to 20 without sigma notation, we have to write
1 + 2 + 3 + 4 + 5 + 6 + 7 + 8 + 9 + 10 + 11 + 12 + 13 + 14 + 15 + 16 + 17 + 18 + 19 + 20.
We could probably skip writing a couple of terms and write
1 + 2 + 3 + 4 + ⋯ + 19 + 20,
which is better, but still cumbersome. With sigma notation, we write this sum as

which is much more compact.


Typically, sigma notation is presented in the form

where ai describes the terms to be added, and the i is called the index. Each term is evaluated, then we sum all
the values, beginning with the value when i = 1 and ending with the value when i = n. For example, an expression

like is interpreted as s2 + s3 + s4 + s5 + s6 + s7. Note that the index is used only to keep track of the terms
to be added; it does not factor into the calculation of the sum itself. The index is therefore called a dummy
variable. We can use any letter we like for the index. Typically, mathematicians use i, j, k, m, and n for indices.
Let’s try a couple of examples of using sigma notation.
Example 5.1
Using Sigma Notation
a. Write in sigma notation and evaluate the sum of terms 3i for i = 1, 2, 3, 4, 5.
b. Write the sum in sigma notation:

Solution

The properties associated with the summation process are given in the following rule.
Rule: Properties of Sigma Notation
Let a1, a2 ,…, an and b1, b2 ,…, bn represent two sequences of terms and let c be a constant. The following
properties hold for all positive integers n and for integers m, with 1 ≤ m ≤ n.

1. (5.1)

2. (5 .2)
3. (5.3)

4. (5.4)

5. (5.5)

Proof
We prove properties 2. and 3. here, and leave proof of the other properties to the Exercises.

A few more formulas for frequently found functions simplify the summation process further. These are shown
in the next rule, for sums and powers of integers, and we use them in the next set of examples.

Rule: Sums and Powers of Integers


Example 5.2
Evaluation Using Sigma Notation
Write using sigma notation and evaluate:
a. The sum of the terms (i − 3)2 for i = 1, 2,…, 200.
b. The sum of the terms (i3–i2) for i = 1, 2, 3, 4, 5, 6.
Solution

Approximating Area
Now that we have the necessary notation, we return to the problem at hand: approximating the area under a curve.
Let f (x) be a continuous, nonnegative function defined on the closed interval [a, b]. We want to approximate the
area A bounded by f(x) above, the x-axis below, the line x = a on the left, and the line x = b on the right (Fig. 5.1).

Fig. 5.1 An area (shaded region) bounded by the curve f(x) at top, the x-axis at bottom, the line x = a to the
left, and the line x = b at right
Definition

We can use this regular partition as the basis of a method for estimating the area under the curve. We next
examine two methods: the left-endpoint approximation and the right-endpoint approximation.
Rule: Left-Endpoint Approximation
On each subinterval [xi − 1, xi] (for i = 1, 2, 3,…, n), construct a rectangle with width Δx and height equal to
f (xi − 1), which is the function value at the left endpoint of the subinterval. Then the area of this rectangle is
f (xi − 1)Δx. Adding the areas of all these rectangles, we get an approximate value for A (Fig. 5.2). We use the
notation Ln to denote that this is a left-endpoint approximation of A using n subintervals.

(5.6)

Fig. 5.2 In the left-endpoint approximation of area under a curve, the height of each rectangle is determined
by the function value at the left of each subinterval
The second method for approximating area under a curve is the right-endpoint approximation. It is almost the
same as the left-endpoint approximation, but now the heights of the rectangles are determined by the function
values at the right of each subinterval.
Rule: Right-Endpoint Approximation
Construct a rectangle on each subinterval [xi − 1, xi], only this time the height of the rectangle is determined
by the function value f (xi) at the right endpoint of the subinterval. Then, the area of each rectangle is f (xi)Δx
and the approximation for A is given by

(5.7)

The notation Rn indicates this is a right-endpoint approximation for A (Fig. 5.3).

Fig. 5.3 In the right-endpoint approximation of area under a curve, the height of each rectangle is determined
by the function value at the right of each subinterval. Note that the right-endpoint approximation differs from
the left-endpoint approximation in Fig. 5.2

The graphs in Fig. 5.4 represent the curve . In graph (a) we divide the region represented by the interval
[0, 3] into six subintervals, each of width 0.5. Thus, Δx = 0.5. We then form six rectangles by drawing vertical
lines perpendicular to xi − 1, the left endpoint of each subinterval. We determine the height of each rectangle by
calculating f (xi − 1) for i = 1, 2, 3, 4, 5, 6. The intervals are [0, 0.5], [0.5, 1], [1, 1.5], [1.5, 2], [2, 2.5], [2.5, 3].
We find the area of each rectangle by multiplying the height by the width. Then, the sum of the rectangular
areas approximates the area between f (x) and the x-axis. When the left endpoints are used to calculate height,
we have a left-endpoint approximation. Thus,
Fig. 5.4 Methods of approximating the area under a curve by using (a) the left endpoints and (b) the right
endpoints
In Fig. 5.4(b), we draw vertical lines perpendicular to xi such that xi is the right endpoint of each subinterval,
and calculate f (xi) for i = 1, 2, 3, 4, 5, 6. We multiply each f (xi) by Δx to find the rectangular areas, and then
add them. This is a right-endpoint approximation of the area under f (x). Thus,

Example 5.3
Approximating the Area Under a Curve
Use both left-endpoint and right-endpoint approximations to approximate the area under the curve
of f (x) = x2 on the interval [0, 2]; use n = 4.
Solution

First, divide the interval [0, 2] into n equal subintervals. Using . This is the
width of each rectangle. The intervals [0, 0.5], [0.5, 1], [1, 1.5], [1.5, 2] are shown in Fig. 5.5.
Using a left-endpoint approximation, the heights are
Fig. 5.5 The graph shows the left-endpoint approximation of the area under f (x) = x2 from 0 to 2
The right-endpoint approximation is shown in Fig. 5.6. The intervals are the same, Δx = 0.5, but
now use the right endpoint to calculate the height of the rectangles. We have

Fig. 5.6 The graph shows the right-endpoint approximation of the area under f (x) = x2 from 0 to 2
The left-endpoint approximation is 1.75; the right-endpoint approximation is 3.75.

Looking at Fig. 5.4 and the graphs in Example 5.3, we can see that when we use a small number of intervals,
neither the left-endpoint approximation nor the right-endpoint approximation is a particularly accurate estimate
of the area under the curve. However, it seems logical that if we increase the number of points in our partition,
our estimate of A will improve. We will have more rectangles, but each rectangle will be thinner, so we will be
able to fit the rectangles to the curve more precisely.
We can demonstrate the improved approximation obtained through smaller intervals with an example. Let’s
explore the idea of increasing n, first in a left-endpoint approximation with four rectangles, then eight rectangles,
and finally 32 rectangles. Then, let’s do the same thing in a right-endpoint approximation, using the same sets
of intervals, of the same curved region. Figure 5.7 shows the area of the region under the curve f (x) = (x − 1)3 + 4
on the interval [0, 2] using a left-endpoint approximation where n = 4. The width of each rectangle is

The area is approximated by the summed areas of the rectangles, or


Fig. 5.7 With a left-endpoint approximation and dividing the region from a to b into four equal intervals, the
area under the curve is approximately equal to the sum of the areas of the rectangles

Figure 5.8 shows the same curve divided into eight subintervals. Comparing the graph with four rectangles in
Figure 5.7 with this graph with eight rectangles, we can see there appears to be less white space under the curve
when n = 8. This white space is area under the curve we are unable to include using our approximation. The
area of the rectangles is

Fig. 5.8 The region under the curve is divided into n = 8 rectangular areas of equal width for a left-endpoint
approximation
The graph in Fig. 5.9 shows the same function with 32 rectangles inscribed under the curve. There appears to
be little white space left. The area occupied by the rectangles is
Fig. 5.9 Here, 32 rectangles are inscribed under the curve for a left-endpoint approximation

5.3 THE DEFINITE INTEGRAL

Definition and Notation


The definite integral generalizes the concept of the area under a curve. We lift the requirements that f (x) be
continuous and nonnegative, and define the definite integral as follows.

Definition
If f (x) is a function defined on an interval [a, b], the definite integral of f from a to b is given by

(5.8)
provided the limit exists. If this limit exists, the function f (x) is said to be integrable on [a, b], or is an
integrable function.

The integral symbol in the previous definition should look familiar. Although the notation for indefinite
integrals may look similar to the notation for a definite integral, they are not the same. A definite integral is a
number. An indefinite integral is a family of functions. Later in this chapter we examine how these concepts are
related. However, close attention should always be paid to notation so we know whether we’re working with a
definite integral or an indefinite integral.
Integral notation goes back to the late seventeenth century and is one of the contributions of Gottfried Wilhelm
Leibniz, who is often considered to be the codiscoverer of calculus, along with Isaac Newton. The integration
symbol ∫ is an elongated S, suggesting sigma or summation. On a definite integral, above and below the
summation symbol are the boundaries of the interval, [a, b]. The numbers a and b are x-values and are called
the limits of integration; specifically, a is the lower limit and b is the upper limit. To clarify, we are using the
word limit in two different ways in the context of the definite integral. First, we talk about the limit of a sum as
n → ∞. Second, the boundaries of the region are called the limits of integration.
We call the function f (x) the integrand, and the dx indicates that f (x) is a function with respect to x, called the
variable of integration. Note that, like the index in a sum, the variable of integration is a dummy variable, and
has no impact on the computation of the integral. We could use any variable we like as the variable of
integration:

Theorem 5.1: Continuous Functions Are Integrable

Example 5.4
Evaluating an Integral Using the Definition

Solution
Evaluating Definite Integrals
Evaluating definite integrals this way can be quite tedious because of the complexity of the calculations. Later in
this chapter we develop techniques for evaluating definite integrals without taking limits of Riemann sums.
However, for now, we can rely on the fact that definite integrals represent the area under the curve, and we can
evaluate definite integrals by using geometric formulas to calculate that area. We do this to confirm that definite
integrals do, indeed, represent areas, so we can then discuss what to do in the case of a curve of a function
dropping below the x-axis.
Area and the Definite Integral
When we defined the definite integral, we lifted the requirement that f(x) be nonnegative. But how do we
interpret “the area under the curve” when f (x) is negative?
Net Signed Area

(shown in Fig. 5.10)

on the interval [0, 2]. Use n = 8 and choose as the left endpoint of each interval. Construct a rectangle

on each subinterval of height and width Δx when is positive, the product represents

the area of the rectangle, as before. When is negative, however, the product represents the
negative of the area of the rectangle. The Riemann sum then becomes

Fig. 5.10 For a function that is partly negative, the Riemann sum is the area of the rectangles above the x-axis
less the area of the rectangles below the x-axis

Taking the limit as n → ∞, the Riemann sum approaches the area between the curve above the x-axis and the
x-axis, less the area between the curve below the x-axis and the x-axis, as shown in Fig. 5.11. Then,
Fig. 5.11 In the limit, the definite integral equals area A1 less area A2, or the net signed area
Notice that net signed area can be positive, negative, or zero. If the area above the x-axis is larger, the net signed
area is positive. If the area below the x-axis is larger, the net signed area is negative. If the areas above and
below the x-axis are equal, the net signed area is zero.
Total Area
One application of the definite integral is finding displacement when given a velocity function. If v(t) represents
the velocity of an object as a function of time, then the area under the curve tells us how far the object is from
its original position. This is a very important application of the definite integral, and we examine it in more
detail later in the chapter. For now, we’re just going to look at some basics to get a feel for how this works by
studying constant velocities.
When velocity is a constant, the area under the curve is just velocity times time. This idea is already very
familiar. If a car travels away from its starting position in a straight line at a speed of 75 mph for 2 hours, then
it is 150 mi away from its original position (Fig. 5.12). Using integral notation, we have
Fig. 5.12 The area under the curve v(t) = 75 tells us how far the car is from its starting point at a given time

In the context of displacement, net signed area allows us to take direction into account. If a car travels straight
north at a speed of 60 mph for 2 hours, it is 120 mi north of its starting position. If the car then turns around and
travels south at a speed of 40 mph for 3 hours, it will be back at it starting position (Fig. 5.13). Again, using
integral notation, we have

In this case the displacement is zero.

Fig. 5.13 The area above the axis and the area below the axis are equal, so the net signed area is zero

Suppose we want to know how far the car travels overall, regardless of direction. In this case, we want to know
the area between the curve and the x-axis, regardless of whether that area is above or below the axis. This is
called the total area.
Graphically, it is easiest to think of calculating total area by adding the areas above the axis and the areas below
the axis (rather than subtracting the areas below the axis, as we did with net signed area). To accomplish this
mathematically, we use the absolute value function. Thus, the total distance traveled by the car is

Bringing these ideas together formally, we state the following definitions.


Definition

Properties of the Definite Integral


The properties of indefinite integrals apply to definite integrals as well. Definite integrals also have properties
that relate to the limits of integration. These properties, along with the rules of integration that we examine later
in this chapter, help us manipulate expressions to evaluate definite integrals.
Rule: Properties of the Definite Integral

1. (5.9)
If the limits of integration are the same, the integral is just a line and contains no area.

2. (5.10)
If the limits are reversed, then place a negative sign in front of the integral.

3. (5.11)
The integral of a sum is the sum of the integrals.

4. (5.12)
The integral of a difference is the difference of the integrals.
5. (5.13)
for constant c. The integral of the product of a constant and a function is equal to the constant multiplied by
the integral of the function.

6. (5.14)
Although this formula normally applies when c is between a and b, the formula holds for all values of a, b,
and c, provided f (x) is integrable on the largest interval.

5.4 INTEGRATION FORMULAS


In this section, we use some basic integration formulas studied previously to solve some key applied problems.
It is important to note that these formulas are presented in terms of indefinite integrals. Although definite and
indefinite integrals are closely related, there are some key differences to keep in mind. A definite integral is
either a number (when the limits of integration are constants) or a single function (when one or both of the
limits of integration are variables). An indefinite integral represents a family of functions, all of which differ
by a constant. As you become more familiar with integration, you will get a feel for when to use definite integrals
and when to use indefinite integrals. You will naturally select the correct approach for a given problem without
thinking too much about it. However, until these concepts are cemented in your mind, think carefully about
whether you need a definite integral or an indefinite integral and make sure you are using the proper notation
based on your choice.
Basic Integration Formulas
Recall the integration formulas given in the table in Antiderivatives and the rule on properties of definite
integrals. Let’s look at a few examples of how to apply these rules.

Rule: Integrals of Even and Odd Functions


5.5 SUBSTITUTION
Theorem 5.7: Substitution with Indefinite Integrals

Theorem 5.8: Substitution with Definite Integrals

Summary
 The properties of indefinite integrals apply to definite integrals as well. Definite integrals also
have properties that relate to the limits of integration.
 A set of points P = {xi} for i = 0, 1, 2,…, n with a = x0 < x1 < x2 < ⋯ < xn = b, which divides
the interval [a, b] into subintervals of the form [x0, x1], [x1, x2],…, [xn −1, xn] is called a partition
of [a, b]. If the subintervals all have the same width, the set of points forms a regular partition
of the interval [a, b].

Keywords
 Total Area: The area between the curve and the x-axis, regardless of whether that area is above
or below the axis.
 Sigma Notation: The Greek capital letter Σ, sigma, is used to express long sums of values in
a compact form.
Self-Assessment Questions
1. Write short note on total area under curve.
2. State and prove theorem on Continuous Functions Are Integrable.
3. Explain Right-Endpoint Approximation.
Unit 6
Current and Resistance
Structure

6.1 Electric Current

6.1.1 Current Density

6.2 Ohm’s Law

6.3 Electrical Energy and Power


Solved Problems
Summary
Self-Assessment Questions
Exercise

The text is adapted by Symbiosis Centre for Distance Learning under a Creative Commons
Attribution-NonCommercial-ShareAlike 4.0 International (CC BY-NC-SA 4.0) as requested
by the work’s creator or licensees. This license is available at
https://creativecommons.org/licenses/by-nc-sa/4.0/
Objectives
After going through this unit, you will be able to
 Explain the concept of Current and Resistance
 To comprehend and apply the Ohm's Law

6.1 ELECTRIC CURRENT


Electric currents are flows of electric charge. Suppose a collection of charges is moving
perpendicular to a surface of area A, as shown in Fig. 6.1.

Fig. 6.1 Charges moving through a cross section

(6.1)
The SI unit of current is the ampere (A), with 1 A = 1 coulomb/sec. Common currents range from
mega-amperes in lightning to nano-amperes in your nerves. In the limit  t  0, the instantaneous
current I may be defined as

(6.2)
Since flow has a direction, we have implicitly introduced a convention that the direction of current
corresponds to the direction in which positive charges are flowing. The flowing charges inside wires
are negatively charged electrons that move in the opposite direction of the current. Electric currents
flow in conductors: solids (metals, semiconductors), liquids (electrolytes, ionized) and gases
(ionized), but the flow is impeded in non- conductors or insulators.
6.1.1 Current Density
To relate current, a macroscopic quantity, to the microscopic motion of the charges, let’s examine
a conductor of cross-sectional area A, as shown in Fig. 6.2.

Fig. 6.2 A microscopic picture of current flowing in a conductor

Let the total current through a surface be written as

(6.3)

(6.4)
The speed vd at which the charge carriers are moving is known as the drift speed. Physically, vd is
the average speed of the charge carriers inside a conductor when an external electric field is
applied. Actually an electron inside the conductor does not travel in a straight line; instead, its path
is rather erratic, as shown in Fig. 6.3.

Fig. 6.3 Motion of an electron in a conductor


From the above equations, the current density can be written as

(6.5)

(6.6)

(6.7)

(6.8)

(6.9)
The current density in Eq. (6.5) becomes

(6.10)
6.2 OHM’S LAW
In many materials, the current density is linearly dependent on the external electric field . Their
relation is usually expressed as

(6.11)
where  is called the conductivity of the material. The above equation is known as the (microscopic)
Ohm’s law. A material that obeys this relation is said to be ohmic; otherwise, the material is non-
ohmic.
Comparing Eq. (6.11) with Eq. (6.10), we see that the conductivity can be expressed as

(6.12)
To obtain a more useful form of Ohm’s law for practical applications, consider a segment of straight
wire of length l and cross-sectional area A, as shown in Fig. 6.4.

Fig. 6.4 A uniform conductor of length l and potential difference V Vb Va

(6.13)
The current density can then be written as

(6.14)
With J  I / A, the potential difference becomes

(6.15)

Where

(6.16)
is the resistance of the conductor. The equation

(6.17)
is the “macroscopic” version of the Ohm’s law. The SI unit of R is the ohm ( , Greek letter Omega),
where

(6.18)
Once again, a material that obeys the above relation is ohmic, and non-ohmic if the relation is not
obeyed. Most metals, with good conductivity and low resistivity, are ohmic. We shall focus mainly
on ohmic materials.

Fig. 6.5 Ohmic vs. Non-ohmic behaviour

The resistivity  of a material is defined as the reciprocal of conductivity,

(6.19)
From the above equations, we see that  can be related to the resistance R of an object by

or

(6.20)
The resistivity of a material actually varies with temperature T. For metals, the variation is linear
over a large range of T:

(6.21)
6.3 ELECTRICAL ENERGY AND POWER
Consider a circuit consisting of a battery and a resistor with resistance R (Fig. 6.6). Let the potential
difference between two points a and b be V Vb Va  0. If a charge q is moved from a through
the battery, its electric potential energy is increased by U  q V. On the other hand, as the charge
moves across the resistor, the potential energy is decreased due to collisions with atoms in the
resistor. If we neglect the internal resistance of the battery and the connecting wires, upon
returning to a the potential energy of q remains unchanged.
Fig. 6.6 A circuit consisting of a battery and a resistor of resistance R
Thus, the rate of energy loss through the resistor is given by

(6.22)
This is precisely the power supplied by the battery. Using V  IR, one may rewrite the above
equation as

(6.23)

Solved Problems
1. Resistivity of a Cable

Solution:
2. Charge at a Junction
Show that the total amount of charge at the junction of the two materials in Fig. 6.7 is

, where I is the current flowing through the junction, and 1 and 2 are the
conductivities for the two materials.

Fig. 6.7 Charge at a junction


Solution:



3. Drift Velocity


4. Resistance of a Truncated Cone
Consider a material of resistivity  in a shape of a truncated cone of altitude h, and radii a and b, for
the right and the left ends, respectively, as shown in the Fig. 6.8.
Fig. 6.8 A truncated Cone

Assuming that the current is distributed uniformly throughout the cross-section of the cone, what is
the resistance between the two ends?
Solution:

Note that if b  a, Eq. (6.19) is reproduced.


5. Resistance of a Hollow Cylinder
Consider a hollow cylinder of length L and inner radius a and outer radius b, as shown in Fig. 6.9.
The material has resistivity .

Fig. 6.9 A hollow cylinder

(a) Suppose a potential difference is applied between the ends of the cylinder and produces a
current flowing parallel to the axis. What is the resistance measured?
(b) If instead the potential difference is applied between the inner and outer surfaces so that
current flows radially outward, what is the resistance measured?
Solution:
Summary
Self-Assessment Questions
1 Two wires A and B of circular cross-section are made of the same metal and have equal
lengths, but the resistance of wire A is four times greater than that of wire B. Find the ratio
of their cross-sectional areas.
2 From the point of view of atomic theory, explain why the resistance of a material increases
as its temperature increases.
3 Two conductors A and B of the same length and radius are connected across the same
potential difference. The resistance of conductor A is twice that of B. To which conductor
is more power delivered?
Exercise

1 A sphere of radius 10 mm that carries a charge of 8 nC  8109 C is whirled in a circle at


the end of an insulated string. The rotation frequency is 100π rad/s.
(a) What is the basic definition of current in terms of charge?
(b) What average current does this rotating charge represent?
(c) What is the average current density over the area traversed by the sphere?
2 A 1500 W radiant heater is constructed to operate at 115 V.
(a) What will be the current in the heater? [Ans. ~10 A]
(b) What is the resistance of the heating coil? [Ans. ~10 Ω]
(c) How many kilocalories are generated in one hour by the heater? (1 Calorie = 4.18 J)
Unit 7
Introduction to Magnetic Fields
Structure
7.1 Introduction
7.2 The Definition of a Magnetic Field
7.3 Magnetic Force on a Current-Carrying Wire
7.4 Torque on a Current Loop
7.4.1 Magnetic Force on a Dipole
7.5 Charged Particles in a Uniform Magnetic Field
7.6 Applications
7.6.1 Velocity Selector
7.6.2 Mass Spectrometer
Summary
Self-Assessment Questions
Exercise

The text is adapted by Symbiosis Centre for Distance Learning under a Creative Commons
Attribution-NonCommercial-ShareAlike 4.0 International (CC BY-NC-SA 4.0) as requested by
the work’s creator or licensees. This license is available at
https://creativecommons.org/licenses/by-nc-sa/4.0/
Objectives
After going through this unit, you will be able to:
 Explain the concept of magnetic field
 To understand the concept of Torque

7.1 INTRODUCTION
We have seen that a charged object produces an electric field at all points in
space. In a similar manner, a bar magnet is a source of a magnetic field . This
can be readily demonstrated by moving a compass near the magnet. The compass
needle will line up along the direction of the magnetic field produced by the
magnet, as depicted in Fig. 7.1.

Fig. 7.1 Magnetic field produced by a bar magnet

Notice that the bar magnet consists of two poles, which are designated as the north
(N) and the south (S). Magnetic fields are strongest at the poles. The magnetic
field lines leave from the north pole and enter the south pole. When holding two
bar magnets close to each other, the like poles will repel each other while the
opposite poles attract (Fig. 7.2).

Fig. 7.2 Magnets attracting and repelling


Unlike electric charges which can be isolated, the two magnetic poles always
come in a pair. When you break the bar magnet, two new bar magnets are
obtained, each with a north pole and a south pole (Fig. 7.3). In other words,
magnetic “monopoles” do not exist in isolation, although they are of theoretical
interest.

Fig. 7.3 Magnetic monopoles do not exist in isolation


How do we define the magnetic field ? In the case of an electric field , we have
already seen that the field is defined as the force per unit charge:

(7.1)
However, due to the absence of magnetic monopoles, must be defined in a
different way.
7.2 THE DEFINITION OF A MAGNETIC FIELD
Fig. 7.4 The direction of the magnetic force

The above observations can be summarized with the following equation:

(7.2)
The above expression can be taken as the working definition of the magnetic field

at a point in space. The magnitude of is given by

(7.3)

(7.4)
The direction of , however, can be altered by the magnetic force, as we shall see
below.
7.3 MAGNETIC FORCE ON A CURRENT-CARRYING WIRE
We have just seen that a charged particle moving through a magnetic field

experiences a magnetic force . Since electric current consists of a collection of


charged particles in motion, when placed in a magnetic field, a current-carrying
wire will also experience a magnetic force.
Consider a long straight wire suspended in the region between the two magnetic
poles. The magnetic field points out the page and is represented with dots (•). It
can be readily demonstrated that when a downward current passes through, the
wire is deflected to the left. However, when the current is upward, the deflection
is rightward, as shown in Fig. 7.5.

Fig. 7.5 Deflection of current-carrying wire by magnetic force

To calculate the force exerted on the wire, consider a segment of wire of length l
and cross-sectional area A, as shown in Fig. 7.6. The magnetic field points into the
page, and is represented with crosses ( X ).

Fig. 7.6 Magnetic force on a conducting wire


(7.5)

For a wire of arbitrary shape, the magnetic force can be obtained by summing over
the forces acting o.n the small segments that make up the wire. Let the differential
segment be denoted as (Fig. 7.7).

Fig. 7.7 Current-carrying wire placed in a magnetic field

The magnetic force acting on the segment is

(7.6)
Thus, the total force is

(7.7)
where a and b represent the endpoints of the wire.
As an example, consider a curved wire carrying a current I in a uniform magnetic

field , as shown in Fig. 7.8.

Fig. 7.8 A curved wire carrying a current I


Using Eq. (7.7), the magnetic force on the wire is given by

(7.8)

Where is the length vector directed from a to b. However, if the wire forms a
closed loop of arbitrary shape (Fig. 7.9), then the force on the loop becomes

(7.9)

Fig. 7.9 A closed loop carrying a current I in a uniform magnetic field


Example 7.1
Magnetic Force on a Semi-Circular Loop
Consider a closed semi-circular loop lying in the xy plane carrying a current I
in the counterclockwise direction, as shown in Fig. 7.10.

Fig. 7.10 Semi-circular loop carrying a current I

A uniform magnetic field pointing in the +y direction is applied. Find the


magnetic force acting on the straight segment and the semicircular arc.
Solution

Let the forces acting on the straight segment and the


semicircular parts, respectively. Using Eq. (7.7) and noting that the length of
the straight segment is 2R, the magnetic force is
7.4 TORQUE ON A CURRENT LOOP
What happens when we place a rectangular loop carrying a current I in the xy
plane and switch on a uniform magnetic field which runs parallel to the
plane of the loop, as shown in Fig. 7.11(a)?

Fig. 7.11 (a) A rectangular current loop placed in a uniform magnetic field. (b)
The magnetic forces acting on sides 2 and 4

From Eq. 7.10, we see the magnetic forces acting on sides 1 and 3 vanish because

the length vectors are parallel and anti-parallel to and


their cross products vanish. On the other hand, the magnetic forces acting on
segments 2 and 4 are non-vanishing:

(7.10)
with pointing out of the page and into the page. Thus, the net force on the
rectangular loop is

(7.11)

as expected. Even though the net force on the loop vanishes, the forces and
will produce a torque which causes the loop to rotate about the y-axis (Fig. 7.12).
The torque with respect to the center of the loop is

(7.12)

(7.13)

Fig. 7.12 Rotation of a rectangular current loop


From Fig. 7.12, the lever arms and can be expressed as:

(7.14)
and the net torque becomes

(7.15)
For a loop consisting of N turns, the magnitude of the toque is

(7.16)

(7.17)

Fig. 7.13 Right-hand rule for determining the direction of

The direction of is the same as the area vector (perpendicular to the plane
of the loop) and is determined by the right-hand rule (Fig. 7.13). The SI unit for
the magnetic dipole moment is ampere-meter2 (A m2) . Using the expression
for , the torque exerted on a current-carrying loop can be rewritten as

(7.18)

The above equation is analogous to , the torque exerted on an electric

dipole moment in the presence of an electric field . Recalling that the

potential energy for an electric dipole is , a similar form is expected


for the magnetic case. The work done by an external agent to rotate the magnetic

dipole from an angle is given by

(7.19)

(7.20)

7.4.1 Magnetic Force on a Dipole


As we have shown above, the force experienced by a current-carrying rectangular
loop (i.e., a magnetic dipole) placed in a uniform magnetic field is zero. What
happens if the magnetic field is non-uniform? In this case, there will be a net force
acting on the dipole.
Consider the situation where a small dipole µ is placed along the symmetric axis
of a bar magnet, as shown in Fig. 7.14.
Fig. 7.14 A magnetic dipole near a bar magnet

(7.21)
where we have used Eq. (7.20). For small x, the external force may be obtained
as

(7.22)

(7.23)

(7.24)
where
(7.25)
is the gradient operator.

Animation 7.1: Torques on a Dipole in a Constant Magnetic Field

“…To understand this point, we have to consider that a


[compass] needle vibrates by gathering upon itself, because of
it magnetic condition and polarity, a certain amount of the lines
of force, which would otherwise traverse the space about it…”
Michael Faraday [1855]

Consider a magnetic dipole in a constant background field. Historically, we note


that Faraday understood the oscillations of a compass needle in exactly the way
we describe here. We show in Fig. 7.15 a magnetic dipole in a “dip needle”
oscillating in the magnetic field of the Earth, at a latitude approximately the same
as that of Boston. The magnetic field of the Earth is predominantly downward and
northward at these Northern latitudes, as the visualization indicates.

Fig. 7.15 A magnetic dipole in the form of a dip needle oscillates in the
magnetic field of the Earth
To explain what is going on in this visualization, suppose that the magnetic dipole
vector is initially along the direction of the earth’s field and rotating clockwise.
As the dipole rotates, the magnetic field lines are compressed and stretched. The
tensions and pressures associated with this field line stretching and compression
results in an electromagnetic torque on the dipole that slows its clockwise rotation.
Eventually the dipole comes to rest. But the counterclockwise torque still exists,
and the dipole then starts to rotate counterclockwise, passing back through being
parallel to the Earth’s field again (where the torque goes to zero), and
overshooting.
As the dipole continues to rotate counterclockwise, the magnetic field lines are
now compressed and stretched in the opposite sense. The electromagnetic torque
has reversed sign, now slowing the dipole in its counterclockwise rotation.
Eventually the dipole will come to rest, start rotating clockwise once more, and
pass back through being parallel to the field, as in the beginning. If there is no
damping in the system, this motion continues indefinitely.

Fig. 7.16 A magnetic dipole in the form of a dip needle rotates oscillates in the
magnetic field of the Earth. We show the currents that produce the earth’s field
in this visualization.

What about the conservation of angular momentum in this situation? Figure 7.16
shows a global picture of the field lines of the dip needle and the field lines of the
Earth, which are generated deep in the core of the Earth. If you examine the
stresses transmitted between the Earth and the dip needle in this visualization, you
can convince yourself that any clockwise torque on the dip needle is accompanied
by a counterclockwise torque on the currents producing the earth’s magnetic field.
Angular momentum is conserved by the exchange of equal and opposite amounts
of angular momentum between the compass and the currents in the Earth’s core.

7.5 CHARGED PARTICLES IN A UNIFORM MAGNETIC FIELD


If a particle of mass m moves in a circle of radius r at a constant speed v, what
acts on the particle is a radial force of magnitude F  mv 2 / r that always points
toward the center and is perpendicular to the velocity of the particle.

In Section 7.2, we have also shown that the magnetic force always points in
the direction perpendicular to the velocity of the charged particle and the

magnetic field . Since can do not work, it can only change the direction of

but not its magnitude. What would happen if a charged particle moves through

a uniform magnetic field with its initial velocity at a right angle to ? For

simplicity, let the charge be +q and the direction of be into the page. It turns

out that will play the role of a centripetal force and the charged particle will
move in a circular path in a counterclockwise direction, as shown in Fig. 7.17.

Fig. 7.17 Path of a charge particle moving in a uniform field with velocity

initially perpendicular to .
With

(7.26)

(7.27)

(7.28)

(7.29)
If the initial velocity of the charged particle has a component parallel to the

magnetic field , instead of a circle, the resulting trajectory will be a helical path,
as shown in Fig. 7.18:

Fig. 7.18 Helical path of a charged particle in an external magnetic field. The

velocity of the particle has a non-zero component along the direction of .


Animation 7.2: Charged Particle Moving in a Uniform Magnetic Field
Figure 7.19 shows a charge moving toward a region where the magnetic field is
vertically upward. When the charge enters the region where the external magnetic
field is non-zero, it is deflected in a direction perpendicular to that field and to its
velocity as it enters the field. This causes the charge to move in an arc that is a
segment of a circle, until the charge exits the region where the external magnetic
field in non-zero. We show in the animation the total magnetic field which is the
sum of the external magnetic field and the magnetic field of the moving charge:

(7.30)

The bulging of that field on the side opposite the direction in which the particle is
pushed is due to the buildup in magnetic pressure on that side. It is this pressure
that causes the charge to move in a circle.

Fig. 7.19 A charged particle moves in a magnetic field that is non-zero over the
pie- shaped region shown. The external field is upward.

Finally, consider momentum conservation. The moving charge in the animation


of Fig. 7.19 changes its direction of motion by 90° over the course of the
animation. How do we conserve momentum in this process? Momentum is
conserved because momentum is transmitted by the field from the moving charge
to the currents that are generating the constant external field. This is plausible
given the field configuration shown in Fig. 7.19. The magnetic field stress, which
pushes the moving charge sideways, is accompanied by a tension pulling the
current source in the opposite direction. To see this, look closely at the field
stresses where the external field lines enter the region where the currents that
produce them are hidden, and remember that the magnetic field acts as if it were
exerting a tension parallel to itself. The momentum loss by the moving charge is
transmitted to the hidden currents producing the constant field in this manner.

7.6 APPLICATIONS

There are many applications involving charged particles moving through a


uniform magnetic field.
7.6.1 Velocity Selector

In the presence of both electric field and magnetic field , the total force on a
charged particle is

(7.30)
This is known as the Lorentz force. By combining the two fields, particles which
move with a certain velocity can be selected. This was the principle used by J. J.
Thomson to measure the charge-to-mass ratio of the electrons. In Fig. 7.20 the
schematic diagram of Thomson’s apparatus is depicted.

Fig. 7.20 Thomson’s apparatus


(7.31)
If the electrons further pass through a region where there exists a downward
uniform electric field, the electrons, being negatively charged, will be deflected
upward. However, if in addition to the electric field, a magnetic field directed
into the page is also.applied, then the electrons will experience an additional

downward magnetic force . When the two forces exactly cancel, the
electrons will move in a straight path. From Eq. 7.30, we see that when the

condition for the cancellation of the two forces is given by , which


implies

(7.32)
In other words, only those particles with speed v = E/B will be able to move in a
straight line. Combining the two equations, we obtain

(7.33)

7.6.2 Mass Spectrometer


Various methods can be used to measure the mass of an atom. One possibility is
through the use of a mass spectrometer. The basic feature of a Bainbridge mass
spectrometer is illustrated in Fig. 7.21. A particle carrying a charge +q is first sent
through a velocity selector.

Fig. 7.21 A Bainbridge mass spectrometer


The applied electric and magnetic fields satisfy the relation E  vB so that the trajectory of

the particle is a straight line. Upon entering a region where a second magnetic field
pointing into the page has been applied, the particle will move in a circular path with radius r
and eventually strike the photographic plate. Using Eq. 7.27, we have

(7.34)
Since v  E / B, the mass of the particle can be written as

(7.35)
Problem-Solving Tips
Solved Problems
1. Rolling Rod
A rod with a mass m and a radius R is mounted on two parallel rails of length a

separated by a distance , as shown in the Fig. 7.22. The rod carries a current I

and rolls without slipping along the rails which are placed in a uniform magnetic
field directed into the page. If the rod is initially at rest, what is its speed as it
leaves the rails?

Fig. 7.22 Rolling rod in uniform magnetic field


Solution:

Using the coordinate system shown on the right,


the magnetic force acting on the rod is given by

The total work done by the magnetic force on the rod as it moves through the
region is .

(7.36)
By the work-energy theorem, W must be equal to the change in kinetic energy:

(7.37)
(7.38)
Thus, the speed of the rod as it leaves the rails is

(7.39)
2. Suspended Conducting Rod
A conducting rod having. a mass density kg/m is suspended by two flexible
wires in a uniform magnetic field which points out of the page, as shown in
Fig. 7.22.

Fig. 7.22 Suspended conducting rod in uniform magnetic field

If the tension on the wires is zero, what are the magnitude and the direction of the
current in the rod?
Solution:

(7.40)
The magnitude of the current can be obtain from

(7.41)
or

(7.42)
3. Charged Particles in Magnetic Field

(7.43)
which yields
(7.44)
The charges move in semicircles, since the magnetic force points radially inward
and provides the source of the centripetal force:

(7.45)
The radius of the circle can be readily obtained as:

(7.46)
1/2
which shows that r is proportional to (m/q) . The mass ratio can then be obtained
from

(7.47)
which gives

(7.48)
4. Bar Magnet in Non-Uniform Magnetic Field
A bar magnet with its north pole up is placed along the symmetric axis below a
horizontal conducting ring carrying current I, as shown in the Fig. 7.23. At the
location of the ring, the magnetic field makes an angle with the vertical. What
is the force on the ring?
Fig. 7.23 A bar magnet approaching a conducting ring
Solution:

The magnetic force acting on a small differential current-carrying element on

the ring is given by , where is the magnetic field due to the bar

magnet. Using cylindrical coordinates as shown in Fig. 7.24, we have

(7.49)
Due to the axial symmetry, the radial component of the force will exactly cancel,
and we are left with the z-component.

Fig. 7.24 Magnetic force acting on the conducting ring


The total force acting on the ring then becomes

(7.50)
The force points in the +z direction and therefore is repulsive.
Summary

 The potential energy of a magnetic dipole placed in a magnetic field is


Self-Assessment Questions
1 Can a charged particle move through a uniform magnetic field without
experiencing any force? Explain.
2 If no work can be done on a charged particle by the magnetic field, how can
the motion of the particle be influenced by the presence of a field?
3 Suppose a charged particle is moving under the influence of both electric and
magnetic fields. How can the effect of the two fields on the motion of the
particle be distinguished?
4 What type of magnetic field can exert a force on a magnetic dipole? Is the force
repulsive or attractive?
5 If a compass needle is placed in a uniform magnetic field, is there a net
magnetic force acting on the needle? Is there a net torque?

Exercise
1. Force Exerted by a Magnetic Field
2. Magnetic Force on a Current Carrying Wire

A square loop of wire, of length on each side, has a mass of 50 g and


pivots about an axis AA' that corresponds to a horizontal side of the square, as
shown in Fig. 7.25. A magnetic field of 500 G, directed vertically downward,
uniformly fills the region in the vicinity of the loop. The loop carries a current I
so that it is in equilibrium at   20.

Fig. 7.25 Magnetic force on a current-carrying square loop


3. Sliding Bar

A conducting bar of length is placed on a frictionless inclined plane which

is tilted at an angle from the horizontal, as shown in Fig. 7.26.

Fig. 7.26 Magnetic force on a conducting bar


A uniform magnetic field is applied in the vertical direction. To prevent the
bar from sliding down, a voltage source is connected to the ends of the bar
with current flowing through. Determine the magnitude and the direction of
the current such that the bar will remain stationary.
Unit 8

Faraday’s Law of Induction


Structure
8.1 Faraday’s Law of Induction
8.1.1 Magnetic Flux
8.1.2 Lenz’s Law
8.2 Motional EMF
8.3 Induced Electric Field
8.4 Generators
8.5 Eddy Currents
Problem-Solving Tips: Faraday’s Law and Lenz’s Law
Solved Problems
Summary
Self-Assessment Questions

The text is adapted by Symbiosis Centre for Distance Learning under a Creative Commons
Attribution-NonCommercial-ShareAlike 4.0 International (CC BY-NC-SA 4.0) as requested
by the work’s creator or licensees. This license is available at
https://creativecommons.org/licenses/by-nc-sa/4.0/
Objectives
After going through this unit, you will be able to:
 To know Faraday's Law of Induction and apply it in day today's life

8.1 FARADAY’S LAW OF INDUCTION


The electric fields and magnetic fields considered up to now have been produced by stationary charges
and moving charges (currents), respectively. Imposing an electric field on a conductor gives rise to a
current which in turn generates a magnetic field. One could then inquire whether or not an electric
field could be produced by a magnetic field. In 1831, Michael Faraday discovered that, by varying
magnetic field with time, an electric field could be generated. The phenomenon is known as
electromagnetic induction. Figure 8.1 illustrates one of Faraday’s experiments.

Fig. 8.1 Electromagnetic induction

Faraday showed that no current is registered in the galvanometer when bar magnet is stationary with
respect to the loop. However, a current is induced in the loop when a relative motion exists between
the bar magnet and the loop. In particular, the galvanometer deflects in one direction as the magnet
approaches the loop, and the opposite direction as it moves away.

Faraday’s experiment demonstrates that an electric current is induced in the loop by changing the
magnetic field. The coil behaves as if it were connected to an emf source. Experimentally it is found
that the induced emf depends on the rate of change of magnetic flux through the coil.
8.1.1 Magnetic Flux
Consider a uniform magnetic field passing through a surface S, as shown in Fig. 8.2 below:

Fig. 8.2 Magnetic flux through a surface

(8.1)

(8.2)

The induced emf  in a coil is proportional to the negative of the rate of change of magnetic
flux:

(8.3)
For a coil that consists of N loops, the total induced emf would be N times as large:

(8.4)
Combining Eqs. (8.3) and (8.1), we obtain, for a spatially uniform field
(8.5)
Thus, we see that an emf may be induced in the following ways:
(i) by varying the magnitude of with time (illustrated in Fig. 8.3.)

Fig. 8.3 Inducing emf by varying the magnetic field strength


(ii) by varying the magnitude of , i.e., the area enclosed by the loop with time
(illustrated in Fig. 8.4.)

Fig. 8.4 Inducing emf by changing the area of the loop


(iii) varying the angle between and the area vector with time (illustrated in Fig. 8.5.)

.
Fig. 8.5 Inducing emf by varying the angle between and
8.1.1 Lenz’s Law
The direction of the induced current is determined by Lenz’s law:

The induced current produces magnetic fields which tend to oppose the change in magnetic
flux that induces such currents.

To illustrate how Lenz’s law works, let’s consider a conducting loop placed in a magnetic field. We
follow the procedure below:

1. Define a positive direction for the area vector .


2. Assuming that is uniform, take the dot product of and . This allows for the
determination of the sign of the magnetic flux  B .

3. Determine the direction of the induced current using the right-hand rule. With your
thumb pointing in the direction of , curl the fingers around the closed loop. The induced
current flows in the same direction as the way your fingers curl if   0 , and the opposite
direction if   0 , as shown in Fig. 8.6.

Fig. 8.6 Determination of the direction of induced current by the right-hand rule
In Fig. 8.7 we illustrate the four possible scenarios of time-varying magnetic flux and show
how Lenz’s law is used to determine the direction of the induced current I.
Fig. 8.7 Direction of the induced current using Lenz’s law
The above situations can be summarized with the following sign convention:

As an example to illustrate how Lenz’s law may be applied, consider the situation where a bar
magnet is moving toward a conducting loop with its north pole down, as shown in Fig. 8.8(a).

With the magnetic field pointing downward and the area vector pointing upward, the

magnetic flux is negative, i.e., , where A is the area of the loop. As the magnet

moves closer to the loop, the magnetic field at a point on the loop increases ,

producing more flux through the plane of the loop. Therefore,


implying a positive induced emf,   0, and the induced current flows in the counterclockwise
direction. The current then sets up an induced magnetic field and produces a positive flux to
counteract the change. The situation described here corresponds to that illustrated in
Fig. 8.7(c).
Alternatively, the direction of the induced current can also be determined from the point of
view of magnetic force. Lenz’s law states that the induced emf must be in the direction that
opposes the change. Therefore, as the bar magnet approaches the loop, it experiences a
repulsive force due to the induced emf. Since like poles repel, the loop must behave as if it
were a bar magnet with its north pole pointing up. Using the right-hand rule, the direction of
the induced current is counterclockwise, as view from above. Figure 8.8(b) illustrates how this
alternative approach is used.

Fig. 8.8 (a) A bar magnet moving toward a current loop. (b) Determination of the direction of
induced current by considering the magnetic force between the bar magnet and the loop
8.2 MOTIONAL EMF
Consider a conducting bar of length l moving through a uniform magnetic field which points
into the page, as shown in Fig, 8.9. Particles with charge q  0 inside experience a magnetic

force which tends to push them upward, leaving negative charges on the lower
end.

Fig. 8.9 A conducting bar moving through a uniform magnetic field


The separation of charge gives rise to an. electric field inside the bar, which in turn

produces a downward electric force . At equilibrium where the two forces cancel, we

have . Between the two ends of the conductor, there exists a potential
difference given by

(8.6)
Since  arises from the motion of the conductor, this potential difference is called the motional
emf. In general, motional emf around a closed conducting loop can be written as

(8.7)

Now suppose the conducting bar moves through a region of uniform magnetic field
(pointing into the page) by sliding along two frictionless conducting rails that are at a distance l
apart and connected together by a resistor with resistance R, as shown in Fig. 8.10.

Fig. 8.10 A conducting bar sliding along two conducting rails

(8.8)
Thus, according to Faraday’s law, the induced emf is

(8.9)
(8.10)
and its direction is counterclockwise, according to Lenz’s law. The equivalent circuit diagram

is shown in Fig. 8.11:


Fig. 8.11 Equivalent circuit diagram for the moving bar

The magnetic force experienced by the bar as it moves to the right is

(8.11)

(8.12)

(8.13)
as required by energy conservation.

(8.14)
or
(8.15)

(8.16)
Thus, we see that the speed decreases exponentially in the absence of an external agent doing
work. In principle, the bar never stops moving. However, one may verify that the total
distance traveled is finite.

8.3 INDUCED ELECTRIC FIELD


The electric potential difference between two points A and B in an electric field can be
written as

(8.17)

Faraday’s law shows that as magnetic flux changes with time, an induced current begins to
flow. What causes the charges to move? It is the induced emf which is the work done per unit
charge. However, since magnetic field can do not work, as we have shown in Unit 7, the work
done on the mobile charges must be electric, and the electric field in this situation cannot be
conservative because the line integral of a conservative field must vanish. Therefore, we

conclude that there is a non-conservative electric field associated with an induced emf:

(8.18)
Combining with Faraday’s law then yields

(8.19)
The above expression implies that a changing magnetic flux will induce a non- conservative
electric field which can vary with time. It is important to distinguish between the induced,
non-conservative electric field and the conservative electric field which arises from electric
charges.

As an example, let’s consider a uniform magnetic field which points into the page and is
confined to a circular region with radius R, as shown in Fig. 8.12. Suppose the magnitude of

increases with time, i.e., dB / dt  0 . Let’s find the induced electric field everywhere due to
the changing magnetic field.

Fig. 8.12 Induced electric field due to changing magnetic flux

(8.20)
Using Eq. (8.19), we have

(8.21)
(8.22)

(8.23)

(8.24)

Fig. 8.13 Induced electric field as a function of r

8.4 GENERATORS
One of the most important applications of Faraday’s law of induction is to generators and
motors. A generator converts mechanical energy into electric energy, while a motor converts
electrical energy into mechanical energy.

Fig. 8.14 (a) A simple generator. (b) The rotating loop as seen from above
Figure 8.14(a) is a simple illustration of a generator. It consists of an N-turn loop rotating in a
magnetic field which is assumed to be uniform. The magnetic flux varies with time, thereby
inducing an emf. From Fig. 8.14(b), we see that the magnetic flux through the loop may be
written as

(8.25)
The rate of change of magnetic flux is

(8.26)
Since there are N turns in the loop, the total induced emf across the two ends of the loop is

(8.27)
If we connect the generator to a circuit which has a resistance R, then the current generated in
the circuit is given by

(8.28)

(8.29)

(8.30)
Thus, the mechanical power supplied to rotate the loop is

(8.31)
Since the dipole moment for the N-turn current loop is
(8.32)
the above expression becomes

(8.33)
As expected, the mechanical power put in is equal to the electrical power output.
8.5 EDDY CURRENTS
We have seen that when a conducting loop moves through a magnetic field, current is induced
as the result of changing magnetic flux. If a solid conductor were used instead of a loop, as
shown in Fig. 8.15, current can also be induced. The induced current appears to be circulating
and is called an eddy current.

Fig. 8.15 Appearance of an eddy current when a solid conductor moves through a magnetic
field

The induced eddy currents also generate a magnetic force that opposes the motion, making it
more difficult to move the conductor across the magnetic field (Fig. 8.16).

Fig. 8.16 Magnetic force arising from the eddy current that opposes the motion of the
conducting slab
Since the conductor has non-vanishing resistance R, Joule heating causes a loss of power by

an amount P   2 / R. Therefore, by increasing the value of R, power loss can be reduced.


One way to increase R is to laminate the conducting slab, or construct the slab by using gluing
together thin strips that are insulated from one another (see Fig. 8.17a). Another way is to
make cuts in the slab, thereby disrupting the conducting path (Fig. 8.17b).

Fig. 8.17 Eddy currents can be reduced by (a) laminating the slab, or (b) making cuts on the
slab
There are important applications of eddy currents. For example, the currents can be used to
suppress unwanted mechanical oscillations. Another application is the magnetic braking
systems in high-speed transit cars.

Problem-Solving Tips: Faraday’s Law and Lenz’s Law


In this unit, we have seen that a changing magnetic flux induces an emf:
3. The sign of the induced emf is the opposite of that of d B / dt . The direction of the induced
current can be found by using Lenz’s law discussed in Section 8.1.2.
Solved Problems
1. Rectangular Loop Near a Wire
An infinite straight wire carries a current I is placed to the left of a rectangular loop of wire
with width w and length l, as shown in the Fig. 8.18.

Fig. 8.18 Rectangular loop near a wire

(a) Determine the magnetic flux through the rectangular loop due to the current I.
(b) Suppose that the current is a function of time with I (t)  a  bt , where a and b are
positive constants. What is the induced emf in the loop and the direction of the
induced current?
Solution
(a) Using Ampere’s law:

(8.34)
the magnetic field due to a current-carrying wire at a distance r away is
(8.35)

(8.36)
Note that we have chosen the area vector to point into the page, so that B  0.
(b) According to Faraday’s law, the induced emf is

(8.37)

The straight wire carrying a current I produces a magnetic flux into the page through the
rectangular loop. By Lenz’s law, the induced current in the loop must be flowing
counterclockwise in order to produce a magnetic field out of the page to counteract the
increase in inward flux.

2. Loop Changing Area


A square loop with length l on each side is placed in a uniform magnetic field pointing into
the page. During a time interval t , the loop is pulled from its two edges and turned into a
rhombus, as shown in the Fig. 8.19. Assuming that the total resistance of the loop is R, find
the average induced current in the loop and its direction.

Fig. 8.19 Conducting loop changing area


Solution:
Using Faraday’s law, we have

(8.38)

(8.39)
Which gives

(8.40)
Thus, the average induced current is

(8.41)

3. Sliding Rod
A conducting rod of length l is free to slide on two parallel conducting bars as in Fig. 8.20.

Fig. 8.20 Sliding rod


In addition, two resistors R1 and R2 are connected across the ends of the bars. There is a
uniform magnetic field pointing into the page. Suppose an external agent pulls the bar to the
left at a constant speed v. Evaluate the following quantities:
(a) The currents through both resistors;
(b) The total power delivered to the resistors;
(c) The applied force needed for the rod to maintain a constant velocity.
Solutions:
(a) The emf induced between the ends of the moving rod is

(8.42)
The currents through the resistors are

(8.43)
Since the flux into the page for the left loop is decreasing, I1 flows clockwise to produce a
magnetic field pointing into the page. On the other hand, the flux into the page for the right
loop is increasing. To compensate the change, according to Lenz’s law, I2 must flow
counterclockwise to produce a magnetic field pointing out of the page.
(b) The total power dissipated in the two resistors is

(8.44)

(8.45)

(8.46)
Summary

Self-Assessment Questions
1. A bar magnet falls through a circular loop, as shown in Figure below
Unit 9
Thermodynamics
Structure
9.1 Introduction

9.2 Thermodynamic Systems

9.3 Work, Heat, and Internal Energy

9.4 First Law of Thermodynamics

9.5 Thermodynamic Processes

9.5.1 Quasi-static and Non-quasi-static Processes

9.5.2 Isothermal Processes

9.5.3 Adiabatic Processes

9.5.4 Cyclic Processes

Summary

Keywords

Self-Assessment Questions

Answers to Check your Questions

Attributes

The text is adapted by Symbiosis Centre for Distance Learning under a Creative Commons Attribution
License 4.0 license (CC BY 4.0) as requested by the work’s creator or licensees. This license is
available at https://creativecommons.org/licenses/by/4.0/

Access for free at https://openstax.org/books/university-physics-volume-2/pages/1-introduction


Objectives

After going through this unit, you will be able to:

 Define a thermodynamic system, its boundary, and its surroundings


 Explain the roles of all the components involved in thermodynamics
 Define thermal equilibrium and thermodynamic temperature
 Describe the work done by a system, heat transfer between objects, and internal energy
change of a system
 State the first law of thermodynamics and explain how it is applied
 Define a thermodynamic process

9.1 INTRODUCTION
Heat is the transfer of energy due to a temperature difference between two systems. Heat describes
the process of converting from one form of energy into another. A car engine, for example, burns
gasoline. Heat is produced when the burned fuel is chemically transformed into
mostly CO2 and H2O, which are gases at the combustion temperature. These gases exert a force on a
piston through a displacement, doing work and converting the piston’s kinetic energy into a variety of
other forms—into the car’s kinetic energy; into electrical energy to run the spark plugs, radio, and
lights; and back into stored energy in the car’s battery.

Energy is conserved in all processes, including those associated with thermodynamic systems. The
roles of heat transfer and internal energy change vary from process to process and affect how work is
done by the system in that process. We will see that the first law of thermodynamics explains that a
change in the internal energy of a system comes from changes in heat or work. Understanding the
laws that govern thermodynamic processes and the relationship between the system and its
surroundings is therefore paramount in gaining scientific knowledge of energy and energy
consumption.

9.2 THERMODYNAMIC SYSTEMS


A thermodynamic system includes anything whose thermodynamic properties are of interest. It is
embedded in its surroundings or environment; it can exchange heat with, and do work on, its
environment through a boundary, which is the imagined wall that separates the system and the
environment. In reality, the immediate surroundings of the system are interacting with it directly and
therefore have a much stronger influence on its behaviour and properties. For example, if we are
studying a car engine, the burning gasoline inside the cylinder of the engine is the thermodynamic
system; the piston, exhaust system, radiator, and air outside form the surroundings of the system. The
boundary then consists of the inner surfaces of the cylinder and piston outside form the surroundings
of the system. The boundary then consists of the inner surfaces of the cylinder and piston.

Access for free at https://openstax.org/books/university-physics-volume-2/pages/1-introduction


Fig. 9.1 (a) A system, which can include any relevant process or value, is self-contained in an area.
The surroundings may also have relevant information; however, the surroundings are important to
study only if the situation is an open system. (b) The burning gasoline in the cylinder of a car engine
is an example of a thermodynamic system

Normally, a system must have some interactions with its surroundings. A system is called an isolated
or closed system if it is completely separated from its environment—for example, a gas that is
surrounded by immovable and thermally insulating walls. In reality, a closed system does not exist
unless the entire universe is treated as the system, or it is used as a model for an actual system that
has minimal interactions with its environment. Most systems are known as an open system, which
can exchange energy and/or matter with its surroundings

Fig. 9.2 (a) This boiling tea kettle is an open thermodynamic system. It transfers heat and matter
(steam) to its surroundings. (b) A pressure cooker is a good approximation to a closed system. A little
steam escapes through the top valve to prevent explosion. (Credit a: modification of work by Gina
Hamilton; Credit b: modification of work by Jane Whitney)

When we examine a thermodynamic system, we ignore the difference in behaviour from place to
place inside the system for a given moment. In other words, we concentrate on the macroscopic
properties of the system, which are the averages of the microscopic properties of all the molecules or
entities in the system. Any thermodynamic system is therefore treated as a continuum that has the

Access for free at https://openstax.org/books/university-physics-volume-2/pages/1-introduction


same behaviour everywhere inside. We assume the system is in equilibrium. You could have, for
example, a temperature gradient across the system. However, when we discuss a thermodynamic
system in this unit, we study those that have uniform properties throughout the system.

Before we can carry out any study on a thermodynamic system, we need a fundamental
characterisation of the system. When we studied a mechanical system, we focused on the forces and
torques on the system, and their balances dictated the mechanical equilibrium of the system. In a
similar way, we should examine the heat transfer between a thermodynamic system and its
environment or between the different parts of the system, and its balance should dictate the thermal
equilibrium of the system. Intuitively, such a balance is reached if the temperature becomes the same
for different objects or parts of the system in thermal contact, and the net heat transfer over time
becomes zero.

Thus, when we say two objects (a thermodynamic system and its environment, for example) are
in thermal equilibrium, we mean that they are at the same temperature, as we discussed
in temperature and heat. Let us consider three objects at temperatures T1 , T2, and T3, respectively.
How do we know whether they are in thermal equilibrium? The governing principle here is the zeroth
law of thermodynamics, as described in Temperature and Heat on temperature and heat:

If object 1 is in thermal equilibrium with objects 2 and 3, respectively, then objects 2 and 3 must also
be in thermal equilibrium.

Mathematically, we can simply write the zeroth law of thermodynamics as

If T1=T2 and T1=T3, then T2=T3. (9.1)


This is the most fundamental way of defining temperature: Two objects must be at the same
temperature thermodynamically if the net heat transfer between them is zero when they are put in
thermal contact and have reached a thermal equilibrium.

The zeroth law of thermodynamics is equally applicable to the different parts of a closed system and
requires that the temperature everywhere inside the system be the same if the system has reached a
thermal equilibrium. To simplify our discussion, we assume the system is uniform with only one type
of material—for example, water in a tank. The measurable properties of the system at least include
its volume, pressure, and temperature. The range of specific relevant variables depends upon the
system. For example, for a stretched rubber band, the relevant variables would be length, tension,
and temperature. The relationship between these three basic properties of the system is called
the equation of state of the system and is written symbolically for a closed system as

f (p,V,T)=0, (9.2)

where V, p, and T are the volume, pressure, and temperature of the system at a given condition.

In principle, this equation of state exists for any thermodynamic system but is not always readily
available. The forms of f(p,V,T)=0 for many materials have been determined either experimentally or
theoretically. We saw an example of an equation of state for an ideal gas, f(p,V,T)=pV−nRT=0.

We have so far introduced several physical properties that are relevant to the thermodynamics of a
thermodynamic system, such as its volume, pressure, and temperature. We can separate these
quantities into two generic categories. The quantity associated with an amount of matter is
an extensive variable, such as the volume and the number of moles. The other properties of a system
are intensive variables, such as the pressure and temperature. An extensive variable doubles its value

Access for free at https://openstax.org/books/university-physics-volume-2/pages/1-introduction


if the amount of matter in the system doubles, provided all the intensive variables remain the same.
For example, the volume or total energy of the system doubles if we double the amount of matter in
the system while holding the temperature and pressure of the system unchanged.

Check your Progress 1


Fill in the Blank.

1. An __________ variable doubles its value if the amount of matter in the system doubles,
provided all the intensive variables remain the same.

9.3 WORK, HEAT, AND INTERNAL ENERGY

We want to understand how work is done by or to a thermodynamic system; how heat is transferred
between a system and its environment; and how the total energy of the system changes under the
influence of the work done and heat transfer.

Work Done by a System

A force created from any source can do work by moving an object through a displacement. Then how
does a thermodynamic system do work? shows a gas confined to a cylinder that has a movable piston
at one end. If the gas expands against the piston, it exerts a force through a distance and does work
on the piston. If the piston compresses the gas as it is moved inward, work is also done—in this case,
on the gas. The work associated with such volume changes can be determined as follows: Let the gas
pressure on the piston face be p. Then the force on the piston due to the gas is pA, where A is the area
of the face. When the piston is pushed outward an infinitesimal distance dx, the magnitude of the
work done by the gas is

dW=F dx=pA dx.

Since the change in volume of the gas is dV=A dx, this becomes

dW=pdV (9.3).

For a finite change in volume from V1 to V2, we can integrate this equation from V1 to V2 to find the
net work:

. (9.4)

Access for free at https://openstax.org/books/university-physics-volume-2/pages/1-introduction


Fig. 9.3 The work done by a confined gas in moving a piston a distance dx is given by dW=F dx=p dV

This integral is only meaningful for a quasi-static process, which means a process that takes place in
infinitesimally small steps, keeping the system at thermal equilibrium. Only then does a well-defined
mathematical relationship (the equation of state) exist between the pressure and volume. This
relationship can be plotted on a pV diagram of pressure versus volume, where the curve is the change
of state. We can approximate such a process as one that occurs slowly, through a series of equilibrium
states. The integral is interpreted graphically as the area under the pV curve (the shaded area of Fig.
9.4). Work done by the gas is positive for expansion and negative for compression.

Fig. 9.4 When a gas expands slowly from V1 to V2, the work done by the system is represented by the
shaded area under the pV curve

Consider the two processes involving an ideal gas that are represented by
paths AC and ABC in Fig. 9.5. The first process is an isothermal expansion, with the volume of the gas
changing its volume from V1 to V2. This isothermal process is represented by the curve between
points A and C. The gas is kept at a constant temperature T by keeping it in thermal equilibrium with
a heat reservoir at that temperature. From Equation 9.4 and the ideal gas law,

Access for free at https://openstax.org/books/university-physics-volume-2/pages/1-introduction


Fig. 9.5 The paths ABC, AC, and ADC represent three different quasi-static transitions between the
equilibrium states A and C

The expansion is isothermal, so T remains constant over the entire process. Since n and R are also
constant, the only variable in the integrand is V, so the work done by an ideal gas in an isothermal
process is

Notice that if V2>V1 (expansion), W is positive, as expected.

The straight lines from A to B and then from B to C represent a different process. Here, a gas at a
pressure p1 first expands isobarically (constant pressure) and quasi-statically from V1 to V2, after which
it cools quasi-statically at the constant volume V2 until its pressure drops to p2. From A to B, the
pressure is constant at p, so the work over this part of the path is

From B to C, there is no change in volume and therefore no work is done. The net work over the
path ABC is then

A comparison of the expressions for the work done by the gas in the two processes of Fig. 9.5 shows
that they are quite different. This illustrates a very important property of thermodynamic work: It
is path dependent. We cannot determine the work done by a system as it goes from one equilibrium
state to another unless we know its thermodynamic path. Different values of the work are associated
with different paths.

Example 9.1
Isothermal Expansion of a van der Waals Gas

Studies of a van der Waals gas require an adjustment to the ideal gas law that takes into
consideration that gas molecules have a definite volume (see The Kinetic Theory of
Gases). One mole of a van der Waals gas has an equation of state

Access for free at https://openstax.org/books/university-physics-volume-2/pages/1-introduction


where a and b are two parameters for a specific gas. Suppose the gas expands
isothermally and quasi-statically from volume V1 to volume V2. How much work is done
by the gas during the expansion?

Strategy

Because the equation of state is given, we can use Equation 9.4 to express the pressure
in terms of V and T. Furthermore, temperature T is a constant under the isothermal
condition, so V becomes the only changing variable under the integral.

Solution

To evaluate this integral, we must express p as a function of V. From the given equation
of state, the gas pressure is

Because T is constant under the isothermal condition, the work done by 1 mol of a van
der Waals gas in expanding from a volume V1 to a volume V2 is thus

Significance

By taking into account the volume of molecules, the expression for work is much more
complex. If, however, we set a=0 and b=0, we see that the expression for work matches
exactly the work done by an isothermal process for one mole of an ideal gas.

ACTIVITY 1

How much work is done by the gas, as given in Fig. 9.5, when it expands quasi-statically along the
path ADC?

Internal Energy

The internal energy Eint of a thermodynamic system is, by definition, the sum of the mechanical
energies of all the molecules or entities in the system. If the kinetic and potential energies of
molecule i are Ki and Ui, respectively, then the internal energy of the system is the average of the total
mechanical energy of all the entities:

(9.5)

Access for free at https://openstax.org/books/university-physics-volume-2/pages/1-introduction


where the summation is over all the molecules of the system, and the bars over K and U indicate
average values. The kinetic energy Ki of an individual molecule includes contributions due to its
rotation and vibration, as well as its translational energy , where vi is the molecule’s speed
measured relative to the center of mass of the system. The potential energy Ui is associated only with
the interactions between molecule i and the other molecules of the system. In fact, neither the
system’s location nor its motion is of any consequence as far as the internal energy is concerned. The
internal energy of the system is not affected by moving it from the basement to the roof of a 100-
story building or by placing it on a moving train.

In an ideal monatomic gas, each molecule is a single atom. Consequently, there is no rotational or
vibrational kinetic energy and . Furthermore, there are no interatomic interactions
(collisions notwithstanding), so Ui=constant, which we set to zero. The internal energy is therefore
due to translational kinetic energy only and

We already know that the average kinetic energy of a molecule in an ideal monatomic gas is

where T is the Kelvin temperature of the gas. Consequently, the average mechanical energy per
molecule of an ideal monatomic gas is also 3kBT/2, that is,

The internal energy is just the number of molecules multiplied by the average mechanical energy per
molecule. Thus for n moles of an ideal monatomic gas,

(9.6)

Notice that the internal energy of a given quantity of an ideal monatomic gas depends on just the
temperature and is completely independent of the pressure and volume of the gas. For other systems,
the internal energy cannot be expressed so simply. However, an increase in internal energy can often
be associated with an increase in temperature.

We know from the zeroth law of thermodynamics that when two systems are placed in thermal
contact, they eventually reach thermal equilibrium, at which point they are at the same temperature.
As an example, suppose we mix two monatomic ideal gases. Now, the energy per molecule of an ideal
monatomic gas is proportional to its temperature. Thus, when the two gases are mixed, the molecules
of the hotter gas must lose energy and the molecules of the colder gas must gain energy. This
continues until thermal equilibrium is reached, at which point, the temperature, and therefore the

Access for free at https://openstax.org/books/university-physics-volume-2/pages/1-introduction


average translational kinetic energy per molecule, is the same for both gases. The approach to
equilibrium for real systems is somewhat more complicated than for an ideal monatomic gas.
Nevertheless, we can still say that energy is exchanged between the systems until their temperatures
are the same.

9.4 FIRST LAW OF THERMODYNAMICS

Now that we have seen how to calculate internal energy, heat, and work done for a thermodynamic
system undergoing change during some process, we can see how these quantities interact to affect
the amount of change that can occur. This interaction is given by the first law of thermodynamics.
British scientist and novelist C. P. Snow (1905–1980) is credited with a joke about the four laws of
thermodynamics. His humorous statement of the first law of thermodynamics is stated “you can’t
win,” or in other words, you cannot get more energy out of a system than you put into it. We will see
in this section how internal energy, heat, and work all play a role in the first law of thermodynamics.

Suppose Q represents the heat exchanged between a system and the environment, and W is the work
done by or on the system. The first law states that the change in internal energy of that system is given
by Q−W. Since added heat increases the internal energy of a system, Q is positive when it is added to
the system and negative when it is removed from the system.

When a gas expands, it does work and its internal energy decreases. Thus, W is positive when work is
done by the system and negative when work is done on the system. This sign convention is
summarized in Table 9.1. The first law of thermodynamics is stated as follows:

FIRST LAW OF THERMODYNAMICS

Associated with every equilibrium state of a system is its internal energy Eint. The change in Eint for any
transition between two equilibrium states is

ΔEint = Q − W (9.7)

where Q and W represent, respectively, the heat exchanged by the system and the work done by or
on the system.

Table 9.1 Thermodynamic Sign Conventions for Heat and Work

Process Convention
Heat added to system Q>0
Heat removed from system Q<0
Work done by system W>0
Work done on system W<0

The first law is a statement of energy conservation. It tells us that a system can exchange energy with
its surroundings by the transmission of heat and by the performance of work. The net energy
exchanged is then equal to the change in the total mechanical energy of the molecules of the system
(i.e., the system’s internal energy). Thus, if a system is isolated, its internal energy must remain
constant.

Although Q and W both depend on the thermodynamic path taken between two equilibrium states,
their difference Q−W does not. Figure 9.6 shows the pV diagram of a system that is making the

Access for free at https://openstax.org/books/university-physics-volume-2/pages/1-introduction


transition from A to B repeatedly along different thermodynamic paths. Along path 1, the system
absorbs heat Q1 and does work W1; along path 2, it absorbs heat Q2 and does work W2, and so on. The
values of Qi and Wi may vary from path to path, but we have

or

That is, the change in the internal energy of the system between A and B is path independent. In the
topic on potential energy and the conservation of energy (to know more about the topic:
https://openstax.org/books/university-physics-volume-1/pages/8-1-potential-energy-of-a-system),
we encountered another path-independent quantity: the change in potential energy between two
arbitrary points in space. This change represents the negative of the work done by a conservative force
between the two points. The potential energy is a function of spatial coordinates, whereas the internal
energy is a function of thermodynamic variables. For example, we might write Eint(T,p) for the internal
energy. Functions such as internal energy and potential energy are known as state functions because
their values depend solely on the state of the system.

Fig. 9.6 Different thermodynamic paths taken by a system in going from state A to state B. For all
transitions, the change in the internal energy of the system ΔEint = Q − W is the same.

Often the first law is used in its differential form, which is

(9.8)

Access for free at https://openstax.org/books/university-physics-volume-2/pages/1-introduction


Here dEint is an infinitesimal change in internal energy when an infinitesimal amount of heat dQ is
exchanged with the system and an infinitesimal amount of work dW is done by (positive in sign) or on
(negative in sign) the system.

Example 9.2
Changes of State and the First Law

During a thermodynamic process, a system moves from state A to state B, it is supplied


with 400 J of heat and does 100 J of work. (a) For this transition, what is the system’s
change in internal energy? (b) If the system then moves from state B back to state A, what
is its change in internal energy? (c) If in moving from A to B along a different
path, W′AB=400 J of work is done on the system, how much heat does it absorb?

Strategy

The first law of thermodynamics relates the internal energy change, work done by the
system, and the heat transferred to the system in a simple equation. The internal energy
is a function of state and is therefore fixed at any given point regardless of how the
system reaches the state.

Solution

a. From the first law, the change in the system’s internal energy is

b. Consider a closed path that passes through the states A and B. Internal energy is
a state function, so ΔEint is zero for a closed path. Thus

and

This yields

c. The change in internal energy is the same for any path, so

and the heat exchanged is

Access for free at https://openstax.org/books/university-physics-volume-2/pages/1-introduction


The negative sign indicates that the system loses heat in this transition.

Significance

When a closed cycle is considered for the first law of thermodynamics, the change in
internal energy around the whole path is equal to zero. If friction were to play a role in
this example, less work would result from this heat added takes into consideration what
happens if friction plays a role.

Notice that in Example 9.2 we did not assume that the transitions were quasi-static. This is because
the first law is not subject to such a restriction. It describes transitions between equilibrium states but
is not concerned with the intermediate states. The system does not have to pass through only
equilibrium states. For example, if a gas in a steel container at a well-defined temperature and
pressure is made to explode by means of a spark, some of the gas may condense, different gas
molecules may combine to form new compounds, and there may be all sorts of turbulence in the
container—but eventually, the system will settle down to a new equilibrium state. This system is
clearly not in equilibrium during its transition; however, its behaviour is still governed by the first law
because the process starts and ends with the system in equilibrium states.

Example 9.3
Polishing a Fitting

A machinist polishes a 0.50-kg copper fitting with a piece of emery cloth for 2.0 min. He
moves the cloth across the fitting at a constant speed of 1.0 m/s by applying a force of
20 N, tangent to the surface of the fitting. (a) What is the total work done on the fitting
by the machinist? (b) What is the increase in the internal energy of the fitting? Assume
that the change in the internal energy of the cloth is negligible and that no heat is
exchanged between the fitting and its environment. (c) What is the increase in the
temperature of the fitting?

Strategy

The machinist’s force over a distance that can be calculated from the speed and time
given is the work done on the system. The work, in turn, increases the internal energy of
the system. This energy can be interpreted as the heat that raises the temperature of the
system via its heat capacity. Be careful with the sign of each quantity.

Solution

a. The power created by a force on an object or the rate at which the machinist

does frictional work on the fitting is . Thus, in an elapsed


time Δt (2.0 min), the work done on the fitting is

Access for free at https://openstax.org/books/university-physics-volume-2/pages/1-introduction


b. By assumption, no heat is exchanged between the fitting and its environment, so
the first law gives for the change in the internal energy of the fitting:

ΔEint = − W = 2.4 × 103 J.

c. Since ΔEint is path independent, the effect of the 2.4 × 103 J of work is the same
as if it were supplied at atmospheric pressure by a transfer of heat. Thus,

and the increase in the temperature of the fitting is

ΔT = 12 °C,

where we have used the value for the specific heat of copper,

Significance

If heat were released, the change in internal energy would be less and cause less of a
temperature change than what was calculated in the problem.

Example 9.4
Vaporizing Water

When 1.00 g of water at 100°C changes from the liquid to the gas phase at atmospheric
pressure, its change in volume is 1.67×10−3m3. (a) How much heat must be added to
vaporize the water? (b) How much work is done by the water against the atmosphere in
its expansion? (c) What is the change in the internal energy of the water?

Strategy

We can first figure out how much heat is needed from the latent heat of vaporization of
the water. From the volume change, we can calculate the work done
from W=pΔV because the pressure is constant. Then, the first law of thermodynamics
provides us with the change in the internal energy.

Solution

a. With Lv representing the latent heat of vaporization, the heat required to


vaporize the water is

b. Since the pressure on the system is constant at 1.00 atm=1.01×105 N/m2, the
work done by the water as it is vaporized is

Access for free at https://openstax.org/books/university-physics-volume-2/pages/1-introduction


c. From the first law, the thermal energy of the water during its vaporization
changes by

Significance

We note that in part (c), we see a change in internal energy, yet there is no change in
temperature. Ideal gases that are not undergoing phase changes have the internal energy
proportional to temperature. Internal energy in general is the sum of all energy in the
system.

ACTIVITY 2

When 1.00 g of ammonia boils at atmospheric pressure and −33.0°C, its volume changes from 1.47
to 1130 cm3. Its heat of vaporization at this pressure is 1.37×106 J/kg. What is the change in the
internal energy of the ammonia when it vaporizes?

Interactive

View this site, https://phet.colorado.edu/en/simulation/legacy/gas-properties to learn about how


the first law of thermodynamics. First, pump some heavy species molecules into the chamber. Then,
play around by doing work (pushing the wall to the right where the person is located) to see how the
internal energy changes (as seen by temperature). Then, look at how heat added changes the internal
energy. Finally, you can set a parameter constant such as temperature and see what happens when
you do work to keep the temperature constant (Note: You might see a change in these variables
initially if you are moving around quickly in the simulation, but ultimately, this value will return to its
equilibrium value).

9.5 THERMODYNAMIC PROCESSES

In solving mechanics problems, we isolate the body under consideration, analyse the external forces
acting on it, and then use Newton’s laws to predict its behaviour. In thermodynamics, we take a similar
approach. We start by identifying the part of the universe we wish to study; it is also known as our
system. Once our system is selected, we determine how the environment, or surroundings, interact
with the system. Finally, with the interaction understood, we study the thermal behaviour of the
system with the help of the laws of thermodynamics.

The thermal behaviour of a system is described in terms of thermodynamic variables. For an ideal gas,
these variables are pressure, volume, temperature, and the number of molecules or moles of the gas.
Different types of systems are generally characterized by different sets of variables. For example, the
thermodynamic variables for a stretched rubber band are tension, length, temperature, and mass.

The state of a system can change as a result of its interaction with the environment. The change in a
system can be fast or slow and large or small. The manner in which a state of a system can change
from an initial state to a final state is called a thermodynamic process. For analytical purposes in
thermodynamics, it is helpful to divide up processes as either quasi-static or non-quasi-static, as we
now explain.

Access for free at https://openstax.org/books/university-physics-volume-2/pages/1-introduction


9.5.1 Quasi-static and Non-quasi-static Processes

A quasi-static process refers to an idealized or imagined process where the change in state is made
infinitesimally slowly so that at each instant, the system can be assumed to be at a thermodynamic
equilibrium with itself and with the environment. For instance, imagine heating 1 kg of water from a
temperature 20°C to 21°C at a constant pressure of 1 atmosphere. To heat the water very slowly, we
may imagine placing the container with water in a large bath that can be slowly heated such that the
temperature of the bath can rise infinitesimally slowly from 20°C to 21°C. If we put 1 kg of water
at 20°C directly into a bath at 21°C, the temperature of the water will rise rapidly to 21°C in a non-
quasi-static way.

Quasi-static processes are done slowly enough that the system remains at thermodynamic equilibrium
at each instant, despite the fact that the system changes over time. The thermodynamic equilibrium
of the system is necessary for the system to have well-defined values of macroscopic properties such
as the temperature and the pressure of the system at each instant of the process. Therefore, quasi-
static processes can be shown as well-defined paths in state space of the system.

Since quasi-static processes cannot be completely realized for any finite change of the system, all
processes in nature are non-quasi-static. Examples of quasi-static and non-quasi-static processes are
shown in Fig. 9.7. Despite the fact that all finite changes must occur essentially non-quasi-statically at
some stage of the change, we can imagine performing infinitely many quasi-static process
corresponding to every quasi-static process. Since quasi-static processes can be analysed analytically,
we mostly study quasi-static processes in this book. We have already seen that in a quasi-static process
the work by a gas is given by pdV.

Fig. 9.7 Quasi-static and non-quasi-static processes between states A and B of a gas. In a quasi-static
process, the path of the process between A and B can be drawn in a state diagram since all the
states that the system goes through are known. In a non-quasi-static process, the states
between A and Bare not known, and hence no path can be drawn. It may follow the dashed line as
shown in the figure or take a very different path

Access for free at https://openstax.org/books/university-physics-volume-2/pages/1-introduction


9.5.2 Isothermal Processes

An isothermal process is a change in the state of the system at a constant temperature. This process
is accomplished by keeping the system in thermal equilibrium with a large heat bath during the
process. Recall that a heat bath is an idealized “infinitely” large system whose temperature does not
change. In practice, the temperature of a finite bath is controlled by either adding or removing a finite
amount of energy as the case may be.

As an illustration of an isothermal process, consider a cylinder of gas with a movable piston immersed
in a large water tank whose temperature is maintained constant. Since the piston is freely movable,
the pressure inside Pin is balanced by the pressure outside Pout by some weights on the piston, as
in Fig. 9.8.

Fig. 9.8 Expanding a system at a constant temperature. Removing weights on the piston leads to an
imbalance of forces on the piston, which causes the piston to move up. As the piston moves up, the
temperature is lowered momentarily, which causes heat to flow from the heat bath to the system.
The energy to move the piston eventually comes from the heat bath

As weights on the piston are removed, an imbalance of forces on the piston develops. The net nonzero
force on the piston would cause the piston to accelerate, resulting in an increase in volume. The
expansion of the gas cools the gas to a lower temperature, which makes it possible for the heat to
enter from the heat bath into the system until the temperature of the gas is reset to the temperature
of the heat bath. If weights are removed in infinitesimal steps, the pressure in the system decreases
infinitesimally slowly. This way, an isothermal process can be conducted quasi-statically. An
isothermal line on a (p, V) diagram is represented by a curved line from starting point A to finishing
point B, as seen in Fig. 9.9. For an ideal gas, an isothermal process is hyperbolic, since for an ideal gas

at constant temperature, .

Access for free at https://openstax.org/books/university-physics-volume-2/pages/1-introduction


Fig. 9.9 An isothermal expansion from a state labeled A to another state labeled B on a pV diagram.
The curve represents the relation between pressure and volume in an ideal gas at constant
temperature

An isothermal process studied in this unit is quasi-statically performed, since to be isothermal


throughout the change of volume, you must be able to state the temperature of the system at each
step, which is possible only if the system is in thermal equilibrium continuously. The system must go
out of equilibrium for the state to change, but for quasi-static processes, we imagine that the process
is conducted in infinitesimal steps such that these departures from equilibrium can be made as brief
and as small as we like.

Other quasi-static processes of interest for gases are isobaric and isochoric processes. An isobaric
process is a process where the pressure of the system does not change, whereas an isochoric
process is a process where the volume of the system does not change.

9.5.3 Adiabatic Processes

In an adiabatic process, the system is insulated from its environment so that although the state of the
system changes, no heat is allowed to enter or leave the system, as seen in Fig. 9.10. An adiabatic
process can be conducted either quasi-statically or non-quasi-statically. When a system expands
adiabatically, it must do work against the outside world, and therefore its energy goes down, which is
reflected in the lowering of the temperature of the system. An adiabatic expansion leads to a lowering
of temperature, and an adiabatic compression leads to an increase of temperature. We discuss
adiabatic expansion again in Adiabatic Processes for an ideal Gas.

Access for free at https://openstax.org/books/university-physics-volume-2/pages/1-introduction


Fig. 9.10 An insulated piston with a hot, compressed gas is released. The piston moves up, the
volume expands, and the pressure and temperature decrease. The internal energy goes into work. If
the expansion occurs within a time frame in which negligible heat can enter the system, then the
process is called adiabatic. Ideally, during an adiabatic process no heat enters or exits the system

9.5.4 Cyclic Processes

We say that a system goes through a cyclic process if the state of the system at the end is same as the
state at the beginning. Therefore, state properties such as temperature, pressure, volume, and
internal energy of the system do not change over a complete cycle:

ΔEint = 0.

When the first law of thermodynamics is applied to a cyclic process, we obtain a simple relation
between heat into the system and the work done by the system over the cycle:

Q=W (cyclic process).

Thermodynamic processes are also distinguished by whether or not they are reversible. A reversible
process is one that can be made to retrace its path by differential changes in the environment. Such a
process must therefore also be quasi-static. Note, however, that a quasi-static process is not
necessarily reversible, since there may be dissipative forces involved. For example, if friction occurred
between the piston and the walls of the cylinder containing the gas, the energy lost to friction would
prevent us from reproducing the original states of the system.

We considered several thermodynamic processes:

1. An isothermal process, during which the system’s temperature remains constant


2. An adiabatic process, during which no heat is transferred to or from the system
3. An isobaric process, during which the system’s pressure does not change
4. An isochoric process, during which the system’s volume does not change

Many other processes also occur that do not fit into any of these four categories.

Access for free at https://openstax.org/books/university-physics-volume-2/pages/1-introduction


Check your Progress 2

State True or False.

1. A reversible process is one that can be made to retrace its path by differential changes in the
environment.

Summary
 Energy is conserved in all processes, including those associated with thermodynamic systems.
The roles of heat transfer and internal energy change vary from process to process and affect
how work is done by the system in that process.
 A thermodynamic system includes anything whose thermodynamic properties are of interest.
It is embedded in its surroundings or environment; it can exchange heat with, and do work
on, its environment through a boundary, which is the imagined wall that separates the system
and the environment.
 A system is called an isolated or closed system if it is completely separated from its
environment—for example, a gas that is surrounded by immovable and thermally insulating
walls.
 Systems are known as an open system, which can exchange energy and/or matter with its
surroundings.
 This integral is only meaningful for a quasi-static process, which means a process that takes
place in infinitesimally small steps, keeping the system at thermal equilibrium.
 The first law is a statement of energy conservation. It tells us that a system can exchange
energy with its surroundings by the transmission of heat and by the performance of work
 The state of a system can change as a result of its interaction with the environment. The
change in a system can be fast or slow and large or small. The manner in which a state of a
system can change from an initial state to a final state is called a thermodynamic process.

Keywords
 Isothermal Processes: An isothermal process is a change in the state of the system at a
constant temperature.
 Adiabatic Processes: In an adiabatic process, the system is insulated from its environment
so that although the state of the system changes, no heat is allowed to enter or leave the
system.

Self-Assessment Questions
1. Describe closed and open systems.
2. Explain Internal energy.
3. Describe first law of thermodynamics.

Answers to Check your Progress

Check your Progress 1

Fill in the Blanks.

1. An extensive variable doubles its value if the amount of matter in the system doubles,
provided all the intensive variables remain the same.

Access for free at https://openstax.org/books/university-physics-volume-2/pages/1-introduction


Check your Progress 2

State True or False.

1. True

Copied from the link: https://cnx.org/contents/eg-XcBxE@16.6:4Mu1W3xW@6/Introduction

Attribute –

Supported by William & Flora Hewlett Foundation, Bill & Melinda Gates Foundation, Michelson
20MM Foundation, Maxfield Foundation, Open Society Foundations, and Rice University. Powered
by OpenStax CNX.
Advanced Placement® and AP® are trademarks registered and/or owned by the College Board, which
was not involved in the production of, and does not endorse, this site.

© 1999-2019, Rice University. Except where otherwise noted, content created


on this site is licensed under a Creative Commons Attribution 4.0 License.

Access for free at https://openstax.org/books/university-physics-volume-2/pages/1-introduction


Unit 10
Heat Transfer Mechanisms
Structure:
10.1 Heat Transfer Modes
10.2 Conduction Heat Transfer
10.3 Convective Heat Transfer
10.4 Temperature Distributions in the Presence of Heat Sources
10.5 Heat Transfer from A Fin
10.6 Transient Heat Transfer (Convective Cooling or Heating)
10.7 Some Considerations in Modeling Complex Physical Processes
10.1 HEAT TRANSFER MODES
Heat transfer processes are classified into three types. The first is conduction, which is defined as transfer of
heat occurring through intervening matter without bulk motion of the matter. Figure shows the process
pictorially. A solid (a block of metal, say) has one surface at a high temperature and one at a lower
temperature. This type of heat conduction can occur, for example, through a turbine blade in a jet engine.
The outside surface, which is exposed to gases from the combustor, is at a higher temperature than the inside
surface, which has cooling air next to it. The level of the wall temperature is critical for a turbine blade.

Fig. 10.1 Conduction heat transfer


The second heat transfer process is convection, or heat transfer due to a flowing fluid. The fluid can be a gas or a
liquid; both have applications in aerospace technology. In convection heat transfer, the heat is moved through bulk
transfer of a non-uniform temperature fluid.
The third process is radiation or transmission of energy through space without the necessary presence of matter.
Radiation is the only method for heat transfer in space. Radiation can be important even in situations in which there is
an intervening medium; a familiar example is the heat transfer from a glowing piece of metal or from a fire.
10.2 CONDUCTION HEAT TRANSFER
We will start by examining conduction heat transfer. We must first determine how to relate the heat transfer to other
properties (either mechanical, thermal, or geometrical). The answer to this is rooted in experiment, but it can be
motivated by considering heat flow along a "bar" between two heat reservoirs at TA, TB as shown in Fig. 10.2. It is

plausible that the heat transfer rate , is a function of the temperature of the two reservoirs, the bar geometry and the
bar properties. (Are there other factors that should be considered? If so, what?). This can be expressed as

(10.1)

(10.2)
Fig. 10.2 Heat transfer along a bar
An argument for the general form of f2 can be made from physical considerations. One requirement, as said,
is f2 = 0 if TA = TB. Using a MacLaurin series expansion, as follows:

(10.3)

(10.4)
We know that f2(0) = 0 . The derivative evaluated at TA = TB (thermal equilibrium) is a measurable property

of the bar. In addition, we know that It also seems reasonable that if we

had two bars of the same area, we would have twice the heat transfer, so that we can postulate that is
proportional to the area. Finally, although the argument is by no means rigorous, experience leads us to

believe that as L increases should get smaller. All of these lead to the generalization (made by Fourier in
1807) that, for the bar, the derivative in equation (10.4) has the form

(10.5)
In equation (10.5), k is a proportionality factor that is a function of the material and the temperature, A is the
cross-sectional area and L is the length of the bar. In the limit for any temperature difference T across a
length x as both L, TA - TB  0, we can say

(10.6)
A more useful quantity to work with is the heat transfer per unit area, defined as

(10.7)

The quantity is called the heat flux and its units are Watts/m2. The expression in (10.6) can be written in
terms of heat flux as

(10.8)

Equation 10.8 is the one-dimensional form of Fourier's law of heat conduction. The proportionality constant
k is called the thermal conductivity. Its units are W / m-K. Thermal conductivity is a well-tabulated property
for a large number of materials. Some values for familiar materials are given in Table 10.1; others can be
found in the references. The thermal conductivity is a function of temperature and the values shown in Table
10.1 are for room temperature.
Table 10.1 Thermal conductivity at room temperature for some metals and non-metals

10.3 CONVECTIVE HEAT TRANSFER


The second type of heat transfer to be examined is convection, where a key problem is determining the boundary
conditions at a surface exposed to a flowing fluid. An example is the wall temperature in a turbine blade because
turbine temperatures are critical as far as creep (and thus blade) life. A view of the problem is given in Fig. 10.3,
which shows a cross-sectional view of a turbine blade. There are three different types of cooling indicated, all meant to
ensure that the metal is kept at a temperature much lower than that of the combustor exit flow in which the turbine
blade operates. In this case, the turbine wall temperature is not known and must be found as part of the solution to the
problem.
Fig. 10.3 Turbine blade heat transfer configuration
To find the turbine wall temperature, we need to analyze convective heat transfer, which means we need to examine
some features of the fluid motion near a surface. The conditions near a surface are illustrated schematically in
Fig. 10.4.

Fig. 10.4 Temperature and velocity distributions near a surface

10.4 TEMPERATURE DISTRIBUTIONS IN THE PRESENCE OF HEAT SOURCES


There are a number of situations in which there are sources of heat in the domain of interest. Examples are:
1) Electrical heaters where electrical energy is converted resistively into heat
2) Nuclear power supplies
3) Propellants where chemical energy is the source
These situations can be analyzed by looking at a model problem of a slab with heat sources  (W/m3) distributed
throughout. We take the outside walls to be at temperature Tw. and we will determine the maximum internal
temperature.
Fig. 10.5 Slab with heat sources (a) overall configuration, (b) elementary slice

With reference to Figure 4.1(b), a steady-state energy balance yields an equation for the heat flux, .

(10.9)

(10.10)
There is a change in heat flux with x due to the presence of the heat sources. The equation for the
temperature is

(10.11)

Equation (10.11) can be integrated once,

(10.12)
and again to give

(10.13)
where a and b are constants of integration. The boundary conditions imposed are T 0 T L Tw.
Substituting these into Equation (10.13) gives

(10.14)

Writing (10.14) in a normalized, non-dimensional fashion gives a form that exhibits in a more useful manner
the way in which the different parameters enter the problem:

(10.15)

This distribution is sketched in Fig. 10.6. It is symmetric about the mid-plane at with half the energy
due to the sources exiting the slab on each side.

Fig. 10.6 Temperature distribution for slab with distributed heat sources
(10.16)

Referring to the temperature distribution of Equation (10.14) gives for the two terms in Equation (10.15),

(10.17)

(10.18)
Evaluating (10.18) at x = 0 and L allows determination of the two constants a and b. This is left as an
exercise for the reader.
10.5 HEAT TRANSFER FROM A FIN

Fins are used in a large number of applications to increase the heat transfer from surfaces. Typically, the fin material
has a high thermal conductivity. The fin is exposed to a flowing fluid, which cools or heats it, with the high thermal
conductivity allowing increased heat being conducted from the wall through the fin. The design of cooling fins is
encountered in many situations and we thus examine heat transfer in a fin as a way of defining some criteria for
design.

A model configuration is shown in Fig. 10.7. The fin is of length L. The other parameters of the problem are
indicated. The fluid has velocity c and temperature T.
Fig. 10.7 Geometry of heat transfer fin

If there is little variation in temperature across the fin, an appropriate model is to say that the temperature within the
fin is a function of x only, T = T(x), and use a quasi-one-dimensional approach. To do this, consider an element, dx, of

the fin as shown in Figure 10.8. There is heat flow of magnitude at the left-hand side and heat flow out of

magnitude at the right hand side. There is also heat transfer around the perimeter on the top,
bottom, and sides of the fin. From a quasi-one-dimensional point of view, this is a situation similar to that with internal
heat sources, but here, for a cooling fin, in each elemental slice of thickness dx there is essentially a heat sink of

magnitude

Fig. 10.8 Element of fin showing heat transfer

(10.19)

From Equation (10.19) we obtain

(10.20)

In terms of the temperature distribution, T(x):

(10.21)

The quantity of interest is the temperature difference (T - T ), and we can change variables to put
Equation (10.21) in terms of this quantity using the substitution

(10.22)
Equation (10.21) can therefore be written as
(10.23)
Equation (10.23) describes the temperature variation along the fin. It is a second order equation and
needs two boundary conditions. The first of these is that the temperature at the end of the fin that joins
the wall is equal to the wall temperature. (Does this sound plausible? Why or why not?)

(10.24)

The second boundary condition is at the other end of the fin. We will assume that the heat transfer from this end
is negligible1. The boundary condition at x = L is

(10.25)

Equation (10.21) can be written in this dimensionless form as

(10.26)
There is one non-dimensional parameter in Equation (10.26), which we will call m and define by

1
Note: We don’t need to make this assumption, and if we were looking at the problem in detail we
would solve it numerically and not worry about whether an analytic solution existed. In the present
case, developing the analytic solution is useful in presenting the structure of the solution as well as the
numbers, so we resort to the mild fiction of no heat transfer at the fin end. We need to assess, after all
is said and done, whether this is appropriate or not.
(10.27)
The equation for the temperature distribution we have obtained is

(10.28)
This second order equation has the solution

(10.29)
(Try it and see). The boundary condition at  = 0

(10.30a)

is The boundary condition at  = 1 is that the temperature gradient is zero or

(10.30b)

(10.31)
This is the solution to Equation (10.26) for a fin with no heat transfer at the tip. In terms of the actual
heat transfer parameters it is written as
(10.32)
The amount of heat removed from wall due to the fin, which is the quantity of interest, can be found by
differentiating the temperature and evaluating the derivative at the wall, x = 0:

(10.33)
or

(10.34)

(10.35)

The solution is plotted in Figure 10.9, which is taken from the book by Lienhard. Several features of the solution
should be noted. First, one does not need fins which have a length such that m is much greater than 3. Second,
the assumption about no heat transfer at the end begins to be inappropriate as m gets smaller than 3, so for very
short fins the simple expression above would not be a good estimate. We will see below how large the error is.
Fig. 10.9 The temperature distribution, tip temperature, and heat flux in a straight one- dimensional fin with the
tip insulated. [Adapted from: Lienhard, A Heat Transfer Textbook, Prentice- Hall publishers]

10.6 TRANSIENT HEAT TRANSFER (CONVECTIVE COOLING OR HEATING)


All the heat transfer problems we have examined have been steady state, but there are often circumstances in
which the transient response to heat transfer is critical. An example is the heating up of gas turbine compressors
as they are brought up to speed during take-off. The disks that hold the blades are large and take a long time to
come to temperature, while the casing is thin and in the path of high velocity compressor flow and thus comes to
temperature rapidly. The result is that the case expands away from the blade tips, sometimes enough to cause
serious difficulties with aerodynamic performance.
To introduce the topic as well as to increase familiarity with modeling of heat transfer problems, we examine a
lumped parameter analysis of an object cooled by a stream. This will allow us to see what the relevant non-
dimensional parameters are and, at least in a quantitative fashion, how more complex heat transfer objects will
behave. We want to view the object as a "lump" described by a single parameter. We need to determine when
this type of analysis would be appropriate. To address this, consider the temperature difference T1 - Tw between
two locations in the object, as shown in Fig. 10.10.

Tw
c 
 T1

T object

Fig. 10.10 Temperature variation in an object cooled by a flowing fluid


If the heat transfer within the body and from the body to the fluid are of the same magnitude,

(10.36)
where L is a relevant length scale, say half the thickness of the object. The ratio of the temperature
difference is

(10.37)
If the Biot number is small the ratio of temperature differences described in Equation (10.37) is also

We can thus say and neglect the temperature non-uniformity


within the object.
The approximation made is to view the object as having a spatially uniform temperature that is a
function of time only. Explicitly, T = T(t). The first law applied to the object is (using the fact that for
solids cp = cv = c)

(10.38)
(10.39)
The initial temperature, T(0), is equal to some known value, which we can call Ti . Using this,

Equation (10.39) can be written in terms of a non-dimensional temperature difference

(10.40)

At time t = 0, this non-dimensional quantity is equal to one. Equation (10.40) is an equation you have

seen before, which has the solution x  aet /  . For the present problem the form is

(10.41)

Fig. 10.11 Voltage change in an R-C circuit


10.7 SOME CONSIDERATIONS IN MODELING COMPLEX PHYSICAL PROCESSES
In Sections 10.5 and 10.6, a number of assumptions were made about the processes that we were attempting to
describe. These are all part of the general approach to modeling of physical systems. The main idea is that for
engineering systems, one almost always cannot compute the process exactly, especially for fluid flow problems.
At some level of detail, one generally needs to model, i.e. to define some plausible behavior for attributes of the
system that will not be computed. Modeling can span an enormous range from the level of our assumption of
uniform temperature within the solid object to a complex model for the small scale turbulent eddies in the flow
past a compressor blade. In carrying out such modeling, it is critical to have a clear idea of just what the
assumptions really mean, as well as the fidelity that we ascribe to the descriptions of actual physical
phenomena, and we thus look at the statements we have made in this context.

Based on this, we said that Tbody is approximately uniform and Tw  Tbody interior. Another aspect is that setting

erases any geometrical detail of the fin cross section. The only place where P and A enter the problem
is in a non-dimensional combination. A third assumption, made in the fin problem, was that the heat transfer at

the far end can be neglected. The solution including this effect, where is an axial Biot number is given as

Equation (10.42). If the quantity is small, you can see that Equation (10.42) reduces to the previous result
(10.31).

(10.42)
and
(10.43)

You might also like