You are on page 1of 260

STUDY GUIDE

Lesson 1
Introduction

ou were on the road, driving, sometime in the early nineties. An NHL


hockey game had started. You switched on your car radio, and immediately
you heard the excited voice of the commentatoralmost screaming
Gretzky is waiting at the blue line...Kurri is at the corner digging out the

Page 2 of 261

puck! Kurri passes the puck...Gretzky is moving away from the blue
line...now he has the puck. Gretzky is moving, faster and faster! He makes
a beautiful move...splits the defensemen...he shoots...he s-c-o-r-e-s!!
If you were a hockey fan, you were quite happy. But you were not
surprised. You expected such wonderful plays from Gretzky, possibly the
greatest hockey player of all time, one whose moves were described by
ecstatic sports writers as poetry in motion. What may surprise you is that
those few lines of the commentator, describing Gretzkys goal, can be used
as starting points for more than half the topics to be discussed in this
course. Consider the statement Gretzky is waiting at the blue line. For the
split second that Gretzky was waiting at the blue line, he was not different
from a book on the tabletop or an electron in equilibrium between two
charged platesa body at rest in the language of a physicist. As Gretzky
moves faster and faster, a physicist observes an accelerating bodylike a
proton in a linear accelerator. When Gretzky shoots the puck, the physicist
thinks of elastic and inelastic collisions, conservation of energy and
momentum, coefficient of friction, etc.
Physics, contrary to popular belief, does not deal with things that are out of
this world. The physical world we live in is the subject matter of physics.
This world is governed by certain laws that apply to everything happening
within its boundariesbe it Gretzkys breakaway goal or one proton
colliding with another. Some of these laws we already know; others are yet
to be found. Our purpose in studying physics is to learn the known laws,
how they were discovered, and how they explain what we observe around
us. With that knowledge, we can attempt to discover those laws that are as
yet unknown.

Objectives
After completing this lesson, you should be able to
1. discuss the relationship between theory and observation in
physics.
2. specify the uncertainty of a measurement, and round it off to the
appropriate number of significant figures.
3. identify the units of measurement used in describing physical
quantities.
4. convert from one system of units to another.

Reading Assignment

Page 3 of 261

Read Sections 1-1, 1-2 and 1-3 in the textbook. They should give you an
idea of the nature and importance of the laws and principles of physics.1
Note: This is a digital textbook (eTextbook). If you havent already done so, access or
download it now through the link on the course home page.

Uncertainty in Measurement

Top

As you have probably realized from the introduction to this unit, the goal of
physics is to discover how our world works and to understand the laws of
nature. However, in science, discovery usually involves making
observations and doing experiments that result in quantitative
measurements. For the measured values (or data) to be meaningful, it is
very important that we recognize the limitations of the instruments used to
obtain them, and estimate the uncertainty of measurement.

Reading Assignment
Read, carefully, Section 1-4 in the textbook.
Shortly, we will solve a problem from the end of the textbook chapter as an
exercise. Throughout the course, we will solve many problems together, as
we apply the material of each lesson. Problem solving is a very important
part of the learning process. Therefore, we encourage you to spend time
solving problems at various degrees of difficulty. This strategy provides a
good test of your understanding of the main concepts, and increases your
efficiency in solving problems.
It is crucially important to your learning that you endeavour to solve the
problems presented in this Study Guide independently, before you read
through the solutions provided. If you reach an impasse and cannot
proceed, then do refer to the solution, but only to the point at which you
can again proceed on your own. Use the solutions to check your work, but
do the work first!

Chapter 1: Problem 2 (page 16)


How many significant figures do each of the following numbers have: (a)
214, (b) 81.60, (c) 7.03, (d) 0.03, (e) 0.0086, (f) 3236, and (g) 8700?
Solution

Page 4 of 261

a. This number has three significant figures, assuming that the last
digit 4 is uncertain. In scientific notation, it is expressed as
2.14102.
b. In this number, we assume that the zero has been included as
the uncertain digit. So there are four significant digits in this
number, and it is expressed as 8.160101 in scientific notation.
c. This number has three significant figures, and it is expressed the
same way in scientific notation.
d. The zeros in this number can be easily removed when the
number is expressed in scientific notation (i.e. 3102). So,
there is only one significant figure.
e. Similarly, this number has only two significant figures, and takes
the form 8.6103 in scientific notation.
f. This number has four significant figures and takes the form
3.236103 in scientific notation.
g. In this number, there is nothing to indicate that the zeros are
significant. So, we assume that they are just place holders and
that 7 is the uncertain digit. The number, then, has two
significant figures and is expressed as 8.7103 in scientific
notation.
End of solution

Reading Assignment
This is a good time to open your Lab Guide and read the introduction to
Experiment 1. The material there should supplement and enhance what you
have studied from the textbook.

Exercise
Answer the questions at the end of Experiment 1 in the Lab Guide. Keep
your answers so that you can include them in your lab report for this
experiment.

Units of Measurement
Christine: How much coffee do you drink, Mary?
Mary: About four.

Page 5 of 261

Top

Christine: Four what?


In this conversation, Marys answer was not complete, because she gave a
quantitative value without specifying the units. There is a big difference
between four extra-large cups per day and four small cups per week!
So, you will not make much sense if you tell a person that the distance
between two points, A and B, is 100. The person will demand to know if
you mean 100 kilometers, 100 feet or maybe 100 angstroms! Physical
measurements are expressed in numbers, but numbers are meaningless
until you clarify what unit of measurement is being used. Furthermore, the
quantity changes depending on the choice of unit. For example, if you
mean that the distance between A and B is 100 miles, a person using
kilometers as the unit of measurement will say that the distance is 160
kilometers. The number 100 changes to 160 because a different unit of
measurement is chosen.
Physicists consider themselves fortunate because all measurable physical
quantities can be expressed in terms of four fundamental units: unit of
length, unit of mass, unit of time and unit of the electric charge. The unit of
velocity, for example, is expressed in terms of the units of length and time
as follows:

To study the physical world, we must settle the units of measurements of


only these four quantities. Once we had many different sets of units, as we
now have different currencies in different countries. The English favoured
the fps system(or the British engineering system), in which length is
measured in feet, mass is measured in a convoluted way as weight
expressed in pounds, and time is measured in seconds. The rest of Europe
at that time worked in the cgs system in which centimeter, gram and
second were used as units of length, mass and time. Now the scientific
world has agreed to use one set of units, called the SI (Systme
International) units, in which length is measured in meters, mass is
measured in kilograms, time is measured in seconds, and electric charge is
measured in coulombs.
In this course, we use SI units. However, you should keep in mind that
there is nothing absolute about this system of units. It has been chosen for
standardization and convenience, not because it is superior to the other
systems in some mystical way. If we wish, we can always change from one
system of units to another, as you will learn in this section, in the same
way that we can change dollars to yen and yen to drachmas.

Page 6 of 261

Reading Assignment
Read Sections 1-5 and 1-6 in the textbook.

Chapter 1: Problem 18 (Page 16)


A typical atom has a diameter of about 1.01010m. (a) What is this in
inches? (b) Approximately how many atoms are there along a 1.0-cm line?
Solution
a. From the tables on the inside of the front cover of the textbook,
we see that the relation between meter and inch is 1 m = 39.37
in. So, the unit conversion proceeds as follows:

b. Since 1 cm = 0.01 m, there are

or one hundred million atoms along a 1.0 cm line.


End of solution
Note: The last two sections in this chapter (1-7 and 1-8) are optional. We encourage
you to read them, but you will not be tested on their contents in the examination.

Reading Assignment
Go over the material in Appendix A (Mathematical Review) to refresh your
memory of the mathematical tools you will need in this course.

Exercise
Solve Problems 1, 9, 11, 13 and 19 at the end of Chapter 1 in the textbook
(page 16). Answers are given at the end of the textbook.

Page 7 of 261

Assignment 1 Questions
Read the section in the Course Manual about Assignments. Below are the
first two questions in Assignment 1. Please keep your answers until you
have answered all the questions of this assignment. Each question is worth
five marks.
1. Perform the following arithmetic operations and write your
answers with the appropriate number of significant figures.
a. 66.1 m + 23.11 m + 0.9 m
b. 2.11 ms 180.1 s
c. (0.036 m/s) 154.1 s
d. 90.00/
e. (23.1 cm3 20.32 cm3 + 19.0 cm3)
2. Perform the following unit conversions, and express the answer in
scientific notation.
a. 0.222 km to mm
b. 10.05 ft to cm
c. 905.2 in3 to liter
d. 24.0 m/s to km/h
e. 5.12 g/cm3 to kg/m3

Before starting the next lesson, test your comprehension of the material covered in
this unit by going back to the objectives. Make certain that you can meet each
objective. If you have difficulties, do not hesitate to contact your tutor to discuss the
problem.

Footnote
1

Note that all references to the textbook are to Physics: Principles with
Applications, 6th ed., by Douglas C. Giancoli (Prentice Hall: Upper Saddle River, NJ:
2005).
Top

STUDY GUIDE

Page 8 of 261

Lesson 2
Displacement, Velocity and Acceleration

he study of motion, technically known as mechanics, forms the basis


of physics. You will spend a large part of your time in this course studying
motions of various types. The study of motion is divided into two parts:
kinematics and dynamics. Kinematics is concerned with describing
motion, and dynamics with analyzing why a body moves in the way it does.
The simplest form of motion is the motion of a body along a straight line,
called motion in one dimension. You will start your study of motion with
the kinematics of this type of motion, that is, the description of the linear
motion. The three important concepts in kinematics are displacement,
velocity and acceleration.
Although we do not usually distinguish between distance and displacement
in our everyday lives, there is a significant difference between the two in
physics. By distance, we mean the total length traveled by a body in
motion. Displacement is defined as the change between a bodys initial
position and its final position; it contains two pieces of information: the
length of the straight line joining the final and initial positions, and the
direction in which the body has been displaced.

In studying motion, we observe how the position of a body changes with


time. Using the two measurable quantitieschange in position and elapsed
timewe define the mathematical concepts of velocity and acceleration.
We then set up a set of equations, called kinematic equations, that relate
change in position and elapsed time to velocity and acceleration. These
simple kinematic equations serve as tools for describing and predicting the
motion of a body.
History does not identify any single person as the discoverer of the
kinematic equations. They are probably the end results of the efforts of
Galileo Galilei (1564-1642) to provide science with a new direction based
on experimental facts and mathematical theories and of the powerful
mathematical techniques developed by Ren Descartes (1596-1649) for
representing geometric forms and physical processes in symbolic languages
that facilitate logical deduction.

Objectives

Page 9 of 261

After completing this lesson, you should be able to


1. explain the importance of the frame of reference in the
description of the motion of a body.
2. provide the mathematical definitions of distance and
displacement, and describe the difference between these
quantities.
3. provide the mathematical definitions of speed and velocity, and
describe the difference between these quantities.
4. explain the difference between average velocity and
instantaneous velocity, and solve problems dealing with both
concepts.
5. give the mathematical definition of acceleration.

Reference Frames and Displacement

Top

If you are walking down the street and a stranger asks you Where is Red
Deer? Your answer, if you live in Edmonton, most probably will be, Red
Deer is about 150 kilometers south of Edmonton. If you are a resident of
Calgary, there is a good chance that you will say, About 145 kilometers
north of here.
Are the two responses the same or different? The question may appear
silly, but let us jump into it anyway. The two answers have provided the
same informationthe location of the city of Red Deeralthough the
quoted numbers and the directions are different in each response. We
understand that the differences, in numbers and directions, are caused by
the fact that the two responses came from two different observers, located
at two different reference points.
The crucial point here is recognizing that we cannot describe the position of
a body in an absolute way. We need a point of reference to define the
position of a body precisely. In our example, we can use either Edmonton
or Calgary as our reference point and locate the precise position of Red
Deer with respect to either of these points. If we know the positional
relationship between Edmonton and Calgarythat Edmonton is 295 km
north of Calgary (or Calgary is 295 km south of Edmonton)we can explain
why the two observers gave two seemingly different answers.

Reading Assignment
Read Section 2-1 in the textbook, and make sure that you understand the

Page 10 of 261

difference between distance and displacement.1


In certain circumstances, such as when a body is moving along a straight
line, the direction of the displacement is always the same and the
magnitude of the displacement equals the distance traveled by the body. In
such situations, we usually do not make an issue of distinguishing between
displacement and distance, since the direction of motion of the body does
not change. However, such situations form only a small portion of this
course.
Quantities, such as displacement, that can be described completely only
when we use both magnitude and direction, are called vectors. Quantities,
such as distance, that can be described completely by magnitude alone are
scalars. As we will see later in the course, when we study the properties of
vectors (when a vector is multiplied or divided by a scalar), we get a new
vector in the same direction as the original one, with a magnitude obtained
by multiplying or dividing the magnitude of the vector by the scalar.

Speed and Velocity


Suppose at a particular time, say at 3:15 PM, we note the position of an
object; for example, a car, a person, or a star far away. After five minutes,
at 3:20 PM, we look again, trying to locate the object. If we find the object
at the same position as before, we conclude that the object has not moved,
and call it a body at rest relative to the observer. On the other hand, if we
find that the object is at a different position, we conclude that the object
has been in motion during this period of time. Studying motion is,
therefore, the process of observing and identifying the position of an object
or a body as time goes by.

Average Velocity
We will begin our study of linear motion with the simplest kinematic
concepts of average speed and average velocity. You are already
familiar with the concept of average speed. If you drive 400 kilometers in
four hours in your newly acquired sports car, you may boast that you
maintained an average speed of 100 km/h. To compute average speed, you
divide the distance traveled by the time taken to cover the distance. The
average velocity, on the other hand, is computed by dividing the
displacement of the object by the time taken to perform the displacement.
EXAMPLE
The position of a ball was recorded at the beginning and the end of a four-

Page 11 of 261

second time interval, as shown in the diagram below. During this time the
ball moved towards the wall and bounced back. Calculate (a) the average
speed and (b) the average velocity of the ball during this interval.

Solution
a. To calculate the average speed, we need to find the total distance
traveled by the ball, which is equal to 3.0 m+5.5 m=8.5 m. So,

b. To calculate the average velocity, we need to find the


displacement of the ball. This is equal to the difference between
the final and initial positions of the ball, and it is equal to
x2x1=5.5 m3.0 m=2.5 m, directed to the right. So,

and also points to the right.


End of solution
Obviously, time and position have the main roles in the study of motion.
Ironically, neither time nor position can be measured in an absolute way.
What we can physically measure are the change in the position of the body
in motion and the time interval over which that change in position takes
place. Using these two measurable quantities, we bring in various
quantitative concepts, such as speed, displacement, velocity, acceleration,
etc., that allow us to describe and predict the motion of the body. The
average velocity is then described by the formula

Page 12 of 261

The symbol , read as delta, is used to indicate a change or difference in


the value of a quantity. For example, if we use t as the symbol for time, t
will denote a change or a difference in time. If you arrive at a bus stop at
3:15 PM and the bus arrives at 3:24 PM, then the difference in time
between your arrival and the arrival of the bus is t=9 min. Note that t is
the change in or difference in time, not the product of and t.
When a body moves along a straight line, we use x to denote the position
of the body, relative to some reference point, at a given time. Now,
suppose the body is at x1 at time t1 and at a later time t2, the position of
the body is x2. Then the body has performed a displacement equal to
x=x2x1
during the time interval
t=t2t1

Reading Assignment
Read Section 2-2 in the textbook, and make sure that you understand the
difference between average speed and average velocity.

Chapter 2: problem 6 (Page 39)


A particle at t1 = 2.0s is at x1 = 3.4 cm and at t2 = 4.5s is at x2 = 8.5
cm. What is its average velocity? Can you calculate its average speed from
these data?
Solution
a. The time interval in this problem is equal to
t=t2t1=(4.5s)(2.0s)=6.5s
To calculate the average velocity, we need, first, to find the
displacement of the particle during this time interval:
x=x2x1=(8.5 cm)(3.4 cm)=5.1 cm
We can now calculate the average velocity of the particle to be

Page 13 of 261

Note that the sign of the velocity is positive, indicating that the direction of
motion is to the right.

b. The average speed of the particle cannot be calculated from these


data. This is because we need to know the path followed by the
particle during its motion in order to calculate the distance
traveled. We do not know if the particle moved from x2 to

x1 directly; maybe it moved some distance forward before


returning back to x1.
End of solution

Chapter 2: problem 8 (Page 39)


According to a rule-of-thumb, every five seconds between a lightning flash
and the following thunder gives the distance to the flash in miles. Assuming
that the flash of light arrives in essentially no time at all, estimate the
speed of sound in m/s from this rule.
Solution: Thunder is the sound of lightning. So, if you happen to be very
close to a lightning flash, you will see the light and hear the thunder almost
instantaneously. However, if you are at a distance from the lightning event,
you will notice a delay in hearing the thunder. This delay occurs because
sound waves travel at a speed much slower than the speed of light.
According to the rule-of-thumb in this problem, sound waves travel a
distance equal to one mile in approximately five seconds. This information
allows us to estimate the speed of sound in air to be

End of solution

Instantaneous Velocity

Top

In your 400-kilometer trip, which you covered in four hours, you said that
your average speed was 100 km/h. However, it is unlikely that the
speedometer of your car was pointing at this speed during the whole trip.

Page 14 of 261

Probably, there were instances when you had to reduce your speed to 80
km/h because of a slow vehicle in front of you, and there were instances
when the speedometer was pointing at 115 km/h. Your speed at a
particular instant in time, as indicated by the speedometer, is called your
instantaneous velocity.

Reading Assignment
Read section 2-3 in the textbook, and make sure that you understand the
difference between average velocity and instantaneous velocity.

Acceleration
Two cars, in adjacent lanes, stopped at a red light at an intersection. When
the traffic light turned green, the car drivers stepped on the gas,
simultaneously, and started moving. Six seconds later, one of the cars
reached a speed of 60 km/h, while the other car reached only 45 km/h. We
conclude that the speed of the first car was changing at a higher rate than
that of the second car. To be more quantitative, the first car was changing
its speed at a rate of 10 km/h every second, while the rate of change of the
second cars speed was 7.5 km/h in each second. This simple example
leads us to a very important concept in kinematics: acceleration.
Acceleration is defined as the rate of change of velocity with respect to
time. Apart from its importance in the description of motion, it is also the
key to understanding what causes motion. It is computed by dividing the
change in the velocity vector by the time interval over which the velocity
change takes place. Thus, acceleration is also a vector quantity. In this
course, we are concerned with motion under constant acceleration; that is,
acceleration in which the magnitude and direction do not change with time.

Reading Assignment
Read Section 2-4 in the textbook to learn the definition of acceleration of a
body in motion as the rate of change of velocity. Note the similarity
between the definitions of average velocity and average acceleration and
between the definitions of instantaneous velocity and instantaneous
acceleration.
EXAMPLE
An advertisement claims that a sports car can accelerate from rest to 100
km/h in 7.6 seconds. What is its acceleration in m/s2?

Page 15 of 261

Solution
In this example, the initial velocity of the car is v0=0 and the final velocity
is v=100km/h The change in the velocity took place over a time interval
equal to t=7.6s. The goal is to find the acceleration a of the sports car in
units of m/s2.
We should first change the unit of the final velocity from km/h to m/s, as
follows:

The appropriate kinematic equation to use is v=v0+at. In this equation, we


have the values of all the variables except the acceleration, which we are
trying to find. By substituting into this equation, we get
27.78 m/s = 0 + a7.6s
which can easily be solved to provide the acceleration of the car

Notice that we rounded to the appropriate number of significant figures in


the last step of the calculation.
End of solution

Chapter 2: problem 13 (Page 39)


An airplane travels 3100 km at a speed of 790 km/h, and then encounters
a tailwind that boosts its speed to 990 km/h for the next 2800 km. What
was the total time for the trip? What was the average speed of the plane
for this trip? [Hint: Think carefully before using Eq. 211d from the
textbook]
Solution
It is obvious that the motion of the plane in this trip has two parts, for
which we can summarize the following data:
Part 1

Part 2

x1= 3100 km

x2= 2800 km

Page 16 of 261

v1= 790 km/h

v2= 990 km/h

t1= ?

t2= ?

The question asks for the total time t and the overall average speed v.
Clearly

t=t1+t2
the sum of the times taken on the individual parts. Equation 2-11d looks
useful, at first, as it gives average speed in terms of the initial and final
speeds. However, it holds only in the case of uniform acceleration, and here
the acceleration is zero throughout most of the trip. The exception is the
very high acceleration when the tailwind boosts the speed by 200 km/h.
We thus decide to proceed using the definition of average speed as total
distance divided by the time taken to travel that distance. We note also
that within each part of the trip, the average and instantaneous speeds are
the same, as there is no acceleration in either portion of the trip; that is,

We can, therefore, write

x1=v1t1 and x2=v2t2


where the individual times for each part of the trip are calculated to be

and

which results in the following total time for the whole trip

t=t1+t2=6.8 h
Since the total distance traveled is equal to

x=x1+x2=5900 km
we can calculate the average speed of the plane over the whole trip to be

Page 17 of 261

End of solution

Chapter 2: problem 17 (Page 39)


A sprinter accelerates from rest to 10.0 m/s in 1.35 s. What is her
acceleration (a) in m/s2 and (b) in km/h2?
Solution
Acceleration is equal to the change in the velocity divided by the elapsed
time. Since the sprinter started from rest, her initial velocity is equal to
zero. So, we have

This acceleration can be expressed in km/h2 as follows

End of solution

Exercise
Solve Problems 3, 7, 14, 15 and 18 at the end of Chapter 2 in the textbook
(pages 39-40).
Answer to 14: average speed = 61 km/h, average velocity = 0.
Answer to 18:

t = 5.2 s

Assignment 1 Questions
3. On her way to school, a child discovered that her loonie is
missing and there is a hole in her pocket. She turned back and
walked 24 m east along the sidewalk. Then, she stopped for 18 s
and decided to head back to school. After walking 11 m west, she

Page 18 of 261

found the loonie. If the child walked at a speed of 0.25 m/s,


calculate,
a. her average velocity while searching for the coin?
b. her average speed while searching for the coin?
4. A car is initially 4.0 km south of a gas station and is moving with
a constant speed of 55 km/h due north. A truck is initially 6.0 km
north of the gas station and is moving with a constant speed of
45 km/h due south. How far are the vehicles from the gas station
when they pass each other?

Before starting the next lesson, test your comprehension of the material covered in
this unit by turning back to the objectives. Make certain that you can meet each
objective. If you have difficulties, do not hesitate to contact your tutor to discuss the
problem.

Footnote
1

Note that all references to the textbook are to Physics: Principles with
Applications, 6th ed., by Douglas C. Giancoli (Prentice Hall: Upper Saddle River, NJ:
2005).
Top

STUDY GUIDE

Lesson 3
Motion at Constant Acceleration

n this lesson, we restrict ourselves to the study of motion of a body


along a straight line with a constant acceleration a. Suppose that at the
initial time t0 the position of a body is x0 and its velocity is v0. If we can

predict the position x and the velocity v of the body at a later time t, we
then say that we are able to describe the motion of the body.
To be able to describe and predict the motion of a body, we need some

Page 19 of 261

mathematical relationships that will connect the quantities x0 and v0 at


time t0, to x and v at time t, and also to the constant acceleration a.

Starting with the definitions of acceleration and velocity, we can establish


three such relationships:

These three kinematic equations enable us to describe and predict the


motion of a body moving along a straight line with constant acceleration.
To make life easier, it is customary to reset the initial time and start from
t0 = 0. The equations then reduce to

It is very important to note that these equations are valid only when the
acceleration is constant. Also, note that the equation for the position x as a
function of time t is that of a parabola. The quadratic formula described in
Appendix A-4 of the textbook is very useful for solving equations of this
sort.
The kinematic equations above enable us to predict the position and
velocity of a body at later times from knowledge of its position, velocity and
acceleration at earlier times. With some manipulations, we can also
compute the position and velocity of a body at earlier times from
knowledge of these values at later times.

Objectives
After completing this lesson, you should be able to
1. write out the kinematic equationsthe mathematical
relationships among displacement, velocity, acceleration and time
that describe the motion of a body.
2. use the kinematic equations to describe and predict the linear
motion of a body under constant acceleration.

Reading Assignment

Page 20 of 261

Read Section 2-5 in the textbook paying close attention to the derivations
of the kinematic equations. Then, read Section 2-6, which provides
examples and useful information on problem solving techniques.1
EXAMPLE
A car accelerates from 35 km/h to 90 km/h in 5.7 seconds.
a. What is its acceleration in m/s2?
b. How far did it travel during this time? Assume constant
acceleration.
Solution
The velocity of the car increases from an initial velocity v0 = 35 km/h to a
final velocity v = 90 km/h in a time interval t = 5.7 s. The goal is to
calculate the acceleration a and the distance traveled by the car during this
period.
We notice that the unit of time in the specification of the time interval is
seconds, while the unit of time in the specification of velocities is hours.
Since we must use the same units throughout, we start by converting the
unit of velocity from km/h to m/s. The results are

v0 = 9.722 m/s
and

v = 25 m/s
a. The next step is to substitute the values given for the unknowns
in the kinematic equation v = v0 + at. We obtain
25 m/s = 9.722 m/s +

a (5.7 s)

and we then calculate the acceleration of the car to be

which we then round to two significant figures and write as

a = 2.7 m/s2
Note: If this is an intermediate result that will be used in further
calculations (as in part b) then it is better to use the value of a as

Page 21 of 261

generated by the calculator, without rounding.

b. To find the distance (x) traveled by the car as it accelerates from


v0 to v we can use the kinematic equation

x = x0 + v0t + at2
such that

x = 0 + 9.722 m/s 5.7 s + 2.68 m/s2 (5.7 s)2 = 99 m


An alternative way to find the distance is to use the kinematic
equation

v2 = v02 + 2ax
such that
(25 m/s)2 = (9.722m/s)2 + 2 (2.68 m/s2)

which will give us the same value for the distance,

End of solution

Chapter 2: problem 26 (Page 40)

Top

In coming to a stop, a car leaves skid marks 92 m long on the highway.


Assuming a deceleration of 7.00 m/s2, estimate the speed of the car just
before braking.
Solution
In this problem, brakes are applied when the car has an initial speed v0.
The final speed is equal to zero and the distance traveled by the car is 92
m. Notice that the acceleration must be negative because the velocity
decreases with time. In summary we have,

Page 22 of 261

To calculate v0 we substitute into the kinematic equation v2 = v02 + 2a(x


x0) such that
0 = v02 + 2 (7.00 m/s2) 92 m.
The speed of the car just before braking is then equal to

End of solution

Chapter 2: Problem 29 (Page 40)


Show that the equation for the stopping distance of a car is ds = v0tR
v02/(2a), where v0 is the initial speed of the car, tR is the drivers reaction
time, and

a is the constant acceleration (and is negative).

Solution
During the reaction time, the car continues to move with constant velocity
v0 until the driver steps on the brakes and starts decelerating. So, if the
drivers reaction time is equal to tR, then the distance traveled by the car
during this period is given by

dR = v0tR
As the driver applies the brakes, the car starts with an initial velocity
v0 and decelerates at a constant rate until it comes to a complete stop (v =
0) after traveling an additional distance d. To compute d, we use the
equation

v2 = v02 + 2ad

Page 23 of 261

and write

Note that, for an object moving in the positive direction, deceleration means negative
acceleration. So, with a negative a, the distance d, above, is a positive quantity. We
can now calculate the total stopping distance of the car as follows

End of solution

Chapter 2: problem 31 (Page 40)


A runner hopes to complete the 10,000-m run in less than 30.0 min. After
exactly 27.0 min, there are still 1100 m to go. The runner must then
accelerate at 0.20 m/s2 for how many seconds to achieve the desired time?

Solution
As shown in the diagram above, the motion of the runner is divided into
three stages. In Stage 1, the runner begins at the starting line and runs
with constant speed covering a distance of 8900 m in 27.0 min (or 1620 s).
Stage 2 begins at point A when the runner realizes that she will not be able
to finish the race on time if she continues at the same speed. So, she
begins accelerating uniformly at a rate of 0.20 m/s2, until she reaches point
B. In Stage 3, the runner covers the remaining distance, moving with
constant speed v3, such that she completes the whole run in 30.0 min.
Now, we will analyze the motion in each stage.

Page 24 of 261

Stage 1: Here the velocity is constant, and it is equal to

Stage 2: The runner here moves with constant acceleration, and the
information we have about her motion in this stage is

Next, we substitute into the kinematic equations

v0t + at2 and we get the following relations:

v = v0 + at and x = x0 +

Stage 3: The motion of the runner in this last stage is simple, because
there is no acceleration. So, as we did in first stage, we can write the
relation
(1100m d) = (180 s t)v3
Now, we substitute the expressions derived (in Stage 2) for v3 and
the equation above, such that
(1100m (5.494

d into

t + 0.10 t2) = (180 t) (5.494 + 0.20 t)

where units are omitted for convenience. After few algebraic manipulations,
the equation above simplifies to the following quadratic equation
0.10 t2 36

t + 111.08 = 0

which, as explained in Appendix A-4, has the following two solutions

Since t must be less than 180 s, then the runner, at point A, must
accelerate for 3.1 s in order to achieve the desired time.

Page 25 of 261

End of solution

Chapter 2: problem 65 (Page 42)

Top

In the design of a rapid transit system, it is necessary to balance the


average speed of a train against the distance between stops. The more
stops there are, the slower the trains average speed. To get an idea of this
problem, calculate the time it takes a train to make a 9.0-km trip in two
situations:
a. the stations at which the trains must stop are 1.8 km apart (a
total of six stations, including those at ends); and
b. the stations are 3.0 km apart (four stations total).
Assume that at each station the train accelerates at a rate of 1.1 m/s2 until
it reaches 90 km/h, then stays at this speed until its brakes are applied for
arrival at the next station, at which time it decelerates at 2.0 m/s2.
Assume it stops at each intermediate station for 20 s.

Solution
To solve this problem, we first analyze the motion of the train between two
consecutive stations. To do that, we divide the motion into three stages,
depending on the acceleration of the train in each stage; we then calculate
the time of each stage. By adding the times corresponding to the three
stages we find the time interval between two stations. From the number of
stations and the stop time we can then calculate the total time of the whole
trip.
Stage 1: In this stage, the train starts from rest and accelerates steadily
until it reaches its maximum velocity of 90 km/h. Before we proceed, let us
first list the information we have about the kinematics of the motion in this
stage:

Page 26 of 261

To calculate the time of this stage, we substitute into the kinematic


equation v = v0 + at such that
25 m/s = 0 + 1.1 m/s2 t1
and we get

The distance traveled by the train during this stage can be calculated using
the kinematic equation v2 = v02 + 2ax such that
(25 m/s)2 = 0 + 2 1.1 m/s2 d1
where

Since, at the moment, we do not have enough information to calculate the


time and distance traveled by the train during the second stage, we skip to
analyzing the motion in the third stage directly.
Stage 3: This stage begins when the driver hits the brakes to slow down,
and ends when the train comes to a complete stop at the next station. As
we did for Stage 1, we start by listing the information we have about the
motion of the train in this last stage.

The time of this stage is calculated by substituting into the kinematic


equation v = v0 + at such that

Page 27 of 261

0 = 25 m/s + (2.0 m/s2) t3


where

The distance traveled by the train while it is trying to stop can be calculated
using the kinematic equation v2 = v02 + 2ax such that
0 = (25 m/s)2 + 2 (2.0 m/s2) d3
where

Stage 2: Note that t1, d1, t3, and d3 do not depend on how far apart the
stations are, but distance d2 and time t2 do depend on the distance
between the stations.
So, when the stations are 1.8 km apart, then

d1 + d2 + d3 = 1.8 km = 1800 m
and the distance d2 covered by the train during Stage 2 is equal to

d2 = 1800 m d1 d3 = 1800 m 284.09 m 156.25 m = 1359.66 m


Since the train moves with constant speed (v = 25 m/s) during this stage,
the time required to cover this distance is calculated using the equation v =
d2/t2, such that

where

So the time between stations is equal to

t = t1 + t2 + t3 = 22.73 s + 54.39 s + 12.5 s = 89.62 s


Since there are six stations, including the first and last, with five station-to-

Page 28 of 261

next-station stretches, and 20-second stops at each of the four stations


between the first and the last, the time for the whole trip is equal to

ttotal = 5 (89.62 s) + 4 (20 s) = 528.1 s


Since this is the last step in the calculation, we will give the answer in
minutes and round it to two significant figures such that

If the distance between stations is increased, the train will still cover the
same distance and take the same time during the acceleration and
deceleration stages. It will, however, take more time and cover greater
distance during Stage 2 of the motion. So, if the stations are 3.0 km apart
we can write

d1 + d2 + d3 = 3.0 km = 3000 m
and the distance covered by the train while moving with constant velocity is
equal to

d2 = 3000 m d1 d3 = 3000 m 284.09 m 156.25 m = 2559.66 m


The time required to cover this distance is equal to

So the travel time between two stations is equal to

t = t1 + t2 + t3 = 22.73 s + 102.39 s + 12.5 s = 137.62 s


Here, we have four stations, including the first and last, with three stationto-next-station stretches, and 20-second stops at each of the two stations
between the first and the last. The time for the whole trip is then equal to

ttotal = 3 (137.62s) + 2 (20s) = 452.86 s


or

ttotal = 7.5 min


End of solution

Chapter 2: problem 67 (Page 43)

Page 29 of 261

In putting, the force with which a golfer strikes a ball is planned so that the
ball will stop within some small distance of the cup, say, 1.0 m long or
short, in case the putt is missed. Accomplishing this from an uphill lie (that
is, putting downhill, see Fig. 2-39 in the textbook) is more difficult than
from a downhill lie. To see why, assume that on a particular green the ball
decelerates constantly at 2.0 m/s2 going downhill, and constantly at
3.0 m/s2 going uphill. Suppose we have an uphill lie 7.0 m from the cup.
Calculate the allowable range of initial velocities we may impart to the ball
so that it stops in the range 1.0 m short to 1.0 m long of the cup. Do the
same for a downhill lie 7.0 m from the cup. What in your results suggests
that the downhill putt is more difficult?
Solution
If a body with an initial velocity v0 experiences deceleration, it will slow
down and eventually come to rest (v = 0) after traveling a distance d.
Using the equation

v2 = v02 + 2ad
we should be able calculate the initial velocity if we know the value of a.
Let us first consider putting downhill. The cup is 7.0 m away from the ball,
and the acceleration is equal to 2.0 m/s2. For the ball to stop 1.0 m short
of the cup, or 6.0 m from the starting point, the initial velocity is calculated
using the equation above such that
0 = v02 + 2 (2.0m/s2) 6.0 m
Solving for v0, we have

Similarly, the ball will stop 1.0 m past the cup, or 8.0 m away from the
starting point, if the initial velocity of the ball is

So, if the golfer can impart an initial velocity between 4.9 m/s and 5.7 m/s,
the ball will stop within one metre of the cup. The golfer must therefore be
accurate to within (5.7 m/s 4.9 m/s) = 0.8 m/s in downhill putting.
In uphill putting, the acceleration of the ball is 3.0 m/s2. By doing
calculations similar to those shown above, we find that the golfer should
impart an initial velocity of between 6.0 m/s and 6.9 m/s in order for the

Page 30 of 261

ball to stop within one metre of the cup. The golfer therefore has to be
accurate within (6.9 m/s 6.0 m/s) = 0.9 m/s in uphill putting. The
smaller range in allowable initial velocities makes the downhill putt a little
more difficult.
End of solution

Exercise
Solve Problems 21, 25, 28 and 32 at the end of Chapter 2 in the textbook
(page 40).
Answer to 28: (a) 113 m, (b) 70 m.
Answer to 32: She should try to stop.

Assignment 1 Questions
5. The driver of an SUV slammed on the breaks when he saw a tree
52 m away blocking the road. The SUV slowed down at a constant
rate of 4.0 m/s2 for 3.6 s before reaching the tree.
a. What was the initial speed of the SUV?
b. With what speed does it strike the tree?
6. A 108 m long train starts from rest (at t = 0) and accelerates
uniformly. At the same time (at t = 0), a car moving with
constant speed in the same direction reaches the back end of the
train. At t = 12 s the car reaches the front of the train. However,
the train continues to speed up and pulls ahead of the car. At
t = 32 s, the car is left behind the train. Determine,
a. the cars speed
b. the trains acceleration
Before starting the next lesson, test your comprehension of the material covered in
this unit by turning back to the objectives. Make certain that you can meet each
objective. If you have difficulties, do not hesitate to contact your tutor to discuss the
problem.

Footnote

Page 31 of 261

1Note

that all references to the textbook are to Physics: Principles with Applications,
6th ed., by Douglas C. Giancoli (Prentice Hall: Upper Saddle River, NJ: 2005).
Top

STUDY GUIDE

Lesson 4
Freely Falling Objects

ll objects near the surface of the earth experience much the same
effect from gravity. They are attracted down, which causes them to
accelerate downward at a constant rate. This constant acceleration is called
the acceleration due to gravity or gravitational acceleration and is
represented by g. To be precise, the value of g is not a true constant, but
changes from place to place, depending on the latitude, longitude and
altitude of the place. However, these changes in value are small enough
that for most practical purposes, we can consider g to be a constant with a
value of 9.80 m/s2.

Objectives
After completing this lesson, you should be able to
1. provide a definition of the phrase acceleration due to gravity.
2. use the kinematic equations to describe and predict the vertical
motion of a body.
3. analyze the motion of an object using graphical methods.
EXAMPLE
A piece of blackboard chalk is released 1.73 m above the floor. How much
time does it take to hit the floor?
Solution
Before we start to solve this example, we will do two things: select an

Page 32 of 261

appropriate coordinate system and extract all


useful information about the motion of the chalk.
The motion described here is an example of a
one-dimensional motion along a vertical straight
line. So, we can set the coordinate system such
that the positive direction of the y-axis points
upward, with the origin at the grounds surface.
In this convention, the upward direction is
positive, and the downward direction is negative.
According to the chosen coordinate system the
initial position of the chalk is y0 = +1.73 m, and
the final position is y = 0. The word released
indicates that the chalk was not given any initial velocity; therefore, v0 =
0. While falling to the ground, the acceleration of the chalk is equal to the
gravitational acceleration. Since it points downward, it must be negative, so
it is written as a = 9.80 m/s2. So, to summarize, we have

Now, we are ready to solve the question. The kinematic equation


containing all the variables above is y = y0 + v0t + at2
By substituting into this equation, we get
0 = 1.73 m + (0)

t + (9.80 m/s2) t2

or
9.80 m/s2 t2 = 1.73 m
The time it takes the chalk to fall to the floor is then calculated to be

End of solution

Exercise

Page 33 of 261

In the example above, calculate the velocity of the chalk, in km/h, just
before it hits the floor.
Answer: 21.0 km/h
The acceleration due to gravity of a freely falling object near the surface of
the Earth always points straight down and has a magnitude equal to 9.80
m/s2. This is true for any object experiencing free fall, regardless of its
shape, size, mass, speed or direction of motion. However, in substituting
the acceleration due to gravity into the kinematic equations we must be
careful about the sign, which depends on the coordinate system used in a
particular situation. If the upward direction is taken to be positive in the
coordinate system used, then the acceleration due to gravity will have a
negative value. So, using such a coordinate system, the gravitational
acceleration is written as

a = g = 9.80 m/s2
Consider an object shot vertically upward, as shown in the diagram below,
from an initial height y0 above the ground. The initial velocity of the object
is positive, since it is moving in the upward direction. The negative
gravitational acceleration of the object, which points downward, acts to
decrease the velocity until it becomes zero at the maximum height.
However, the object continues to accelerate downward, even when its
velocity is zero at the top of the flight. So, the velocity will keep changing,
causing the object to start moving down with increasing speed until it hits
the ground with a speed equal to the initial speed, assuming that object
returned back to its initial position. If the object moves so that it goes
higher than its release point (i.e., if the displacement y is positive), the
final speed will be less than the initial speed. Conversely, if the object falls
to below the point at which it started, it must have a greater final speed.

Reading Assignment
Read Section 2-7 in the textbook, paying close attention to the examples.
It is always a good strategy to try to solve the examples on your own,
before looking at the solution provided.1

Chapter 2: problem 35 (Page 40)


Estimate:
a. how long it took King Kong to fall straight down from the top of

Page 34 of 261

Top

the
Empire
State
Building
(380 m
high); and
b. his
velocity
just
before
landing.
Solution
If King Kong
simply fell,
then his initial
velocity would be zero, and he would experience a free fall under the
gravitational acceleration. Here is a summary of the information we have
about King Kongs motion:

a. The information above suggests that the appropriate kinematic


equation for calculating the time is y = y0 + v0t + at2. In this
equation, all the parameters are known except the time, which
we are trying to find. By substitution, we get
0 = 380 m + (0)

t + (9.80 m/s2) t2

and

b. To find King Kongs velocity when he hit the ground, we will use
the time calculated above and substitute into the equation v = v0
+

at, such that

Page 35 of 261

v = 0 + 9.80m/s2 8.8s = 86 m/s


An alternative method for calculating the final velocity is
substituting into the kinematic equation v2 = v02 + 2a(y

y0) such that


v2 = 0 + 2 (9.80 m/s2) (0 380 m)
which leads to the same result as above:

Therefore, King Kong spent 8.8 s in the air before hitting the
ground with a velocity 86 m/s.
End of solution

Chapter 2: problem 36 (Page 40)


A baseball is hit nearly straight up into the air with a speed of 22 m/s. (a)
How high does it go? (b) How long is it in the air?
Solution
When it is hit, the ball starts moving in the upward direction with an initial
velocity equal to 22 m/s. On its way up, the ball slows down due to
gravitational acceleration, until it stops (momentarily) at the maximum
height. A summary of the information we have about the upward motion of
the baseball is given below.

To calculate the upward distance (i.e., the maximum height) traveled by


the ball, we use the kinematic equation v2 = v02 + 2a(y y0) By
substitution, we get
0 = (22 m/s)2 + 2 (9.80 m/s2)

where the maximum height reached by the ball is calculated to be

Page 36 of 261

From the symmetry of the problem, we can argue that the ball takes an
equal time for its upward trip and its downward trip. Therefore, the ball
spends in the air twice the time it takes to reach the maximum height. So,
by substituting into the equation v = v0 + at such that
0 = 22 m/s + (9.80 m/s2)

we can calculate the time t, which the ball takes to reach its maximum
height, as follows:

So the ball spends a time equal to

tair = 2t = 4.5 s
in the air.
Another method for solving this problem would involve calculating tair,
directly, using the kinematic equation

y = y0 + v0t + at2 substituting (y = y0 = 0)


This is left for you as an exercise.
End of solution

Chapter 2: problem 45 (Page 41)


A rock is dropped from a sea cliff, and the sound of it striking the ocean is
heard 3.2 s later. If the speed of sound is 340 m/s, how high is the cliff?
Solution
Suppose that the height of the cliff above sea level is h. If tr is the time
required by the rock to reach the sea surface, then we can summarize the
information we have about the motion of the rock as follows:

Page 37 of 261

We then substitute into the kinematic equation

y = y0 + v0t + at2 such

that
0=

h + 0 + (9.80 m/s2) tr2

where we get the following relation between

h and tr

h = (4.90 m/s2) tr2

The sound of the splash travels upwards with a constant velocity vs = 340
m/s. If ts is the time taken by the sound waves to travel up to the top of
the cliff, then we get the following relation between

h and ts

h = (340 m/s) ts

Since the sound waves arrived 3.2 s after the rock was dropped, we can
write the following relation between tr and ts

tr + ts = 3.2 s

Page 38 of 261

We now have three equations in three unknowns, which we need to solve


simultaneously to find h. The rest of the solution to this problem is, then,
pure algebra. One way to proceed is by combining equations (1) and (2) as
follows
(4.90 m/s2) tr2 = (340 m/s) ts
and then use equation (3) to substitute for ts such that
(4.90 m/s2) tr2 = (340 m/s) (3.2 s tr)
This substitution leads to the following quadratic equation in tr
4.9 tr2 + 340 tr 1088 = 0
where we have omitted the units for convenience. Recalling that the roots
of the quadratic equation ax2 + bx + c = 0 are

we get

We ignore the negative value, since it does not have any physical
significance, and take tr = 3.06 s. If we now substitute tr into equation (1),
we can calculate the height of the cliff to be

h = 4.90 m/s2 (3.06s)2 = 46 m


End of solution

Chapter 2: problem 46 (Page 41)


Suppose you adjust your garden hose nozzle for a hard stream of water.
You point the nozzle vertically upward at a height of 1.5 m above the
ground (Fig. 2-33 in the textbook). When you quickly move the nozzle
away from the vertical, you hear the water striking the ground next to you
for another 2.0 s. What is the water speed as it leaves the nozzle?
Solution

Page 39 of 261

To solve this problem, we will study the


motion of the last spurt of water that
leaves the nozzle before it is moved from
the vertical position. This last spurt of
water shoots up from an initial position
1.5 m above the ground and undergoes a
free-fall during its journey up to the
maximum height and then down all the
way to the ground. In summary, the
information we have about this motion is

The solution to this problem becomes


straightforward if we substitute into the kinematic equation
at2 such that

y = y0 + v0t +

0 = 1.5 m + v0 2.0s + (9.80m/s2) (2.0s)2


Solving for v0 we find the following value for the (initial) velocity of the
water as it leaves the nozzle:

End of solution

Chapter 2: problem 69 (Page 43)


A stone is dropped from the roof of a high building. A second stone is
dropped 1.50 s later. How far apart are the stones when the second one
has reached a speed of 12.0 m/s?
Solution
In this problem, it is more convenient to use a coordinate system having
the origin at the roof of the building and with positive direction pointing
down. In this case, the gravitational acceleration is positive.
We begin by a summary of the information we have about the motion of

Page 40 of 261

each stone from the time of release until the time when the second stone
reached a speed of 12.0 m/s:
1st Stone
y0 = 0

2nd Stone
y0 = 0

y = y1

y = y2

v0 = 0

v0 = 0

v=?
v = 12 m/s
a = 9.80 m/s2 a = 9.80 m/s2
t = t2 + 1.5 s t = t2 = ?
Since there is more information available about the motion of the second
stone, we will start by finding t2, which is the time taken by the second
stone to reach a speed of 12.0 m/s. This can be easily calculated by
substituting into the kinematic equation v = v0 + at such that
12 m/s = 0 + 9.80 m/s2 t2
where we have

Now, we can calculate the downward distance moved by the second stone
during this time by substituting into the equation y = y0 + v0t + at2 such
that

y2 = 0 + 0 + 9.80 m/s2 (1.22s)2 = 7.3 m


Since the first stone started moving 1.5 s earlier, then it will have fallen a
distance of

y1 = 0 + 0 + (9.80 m/s2) (1.5 s + 1.22s)2 = 36.3 m


below the roof of the building.
The separation between the two stones is then equal to

y1 y2 = 36.3 m 7.3 m = 29 m
End of solution

Page 41 of 261

Exercise
Solve Problems 33, 39, 44, 47 and 83 at the end of Chapter 2 in the
textbook (pages 40-44).
Answer to 44: 2.1 m.

Graphical Representation of Linear Motion

Top

Suppose that we observe a body moving in a straight Time Position


line from left to right. We measure the position x of
t (s) x (m)
the body at different times t, and tabulate
5
10
measurements as shown. These observations can be
summarized in a graph form. We draw two
10
20
perpendicular axes corresponding to the two
15
30
measured quantities; the horizontal axis for the time
and the vertical axis for the position, as shown in the
20
40
diagram below. Point A(5, 10) on the graph says that
at time t = 5 s the body was at the position x = 10 m. Similarly, points B
(10, 20), C(15, 30) and D(20, 40) indicate that at 10 s, 15 s and 20 s from
the start, the body was observed at 20 m, 30 m and 40 m from the
reference point. In general, we represent a point on a graph as P(t, x), to
indicate that at time t the body is at the position x.
We can draw a smooth line
joining the points A, B, C and D,
as shown in the diagram. The
resulting graph provides a
representation of the relationship
between position x and time t of
the motion of the body. Such a
graph is known as the timeposition graph of a body in
motion. The time-position graph
in this example is a straight line.
If the relationship between two variables, x and t in our case, can be
represented by a straight line, mathematicians say that the relationship is
linear. A linear relationship between x and t indicates that the body moves
with a constant velocity; that is, the body does not accelerate with time.
On the other hand, a nonlinear relationship occurs when a body has a
nonzero acceleration; that is, when the velocity of the body changes with
time. The most important nonlinear motion studied in this course occurs

Page 42 of 261

when the acceleration of the body is constant. The position of the body as a
function of time, in this case, is represented by a special curve called a
parabola (see Figure 2-21b in the textbook).
Returning to our example, in moving from A to B, the body travels a
distance x = (20 m 10 m) = 10 m, in a period of time t = (10 s 5 s)
= 5 s. Therefore, the average velocity of the body as it makes the
transition from A to B is

Between B and C, the change in position is x = (30 m 20 m) = 10 m, in


the time interval t = (15 s 10 s) = 5 s. The average velocity as the
body makes the transition from B to C is then given by

The average velocity of the body does not change as the body moves from
A to B to C. As we shall see, the time-position graph of motion of a body
moving with a constant velocity is always a straight line.
Let us generalize the analysis we have just completed. Consider the timeposition graph of motion of a body shown in the figure below.

We are interested in the two points A(t1, x1) and B(t2, x2). Let v be the
average velocity of the body as it makes the transition from A to B. As the
body has moved through a distance x = x2 x1 in time t = t2 t1, the

Page 43 of 261

average velocity is

If we complete the right triangle ABC, as shown, we see that


x = x2 x1 = BC
and
t = t2 t1 = AC
so that

The ratio BC/AC is called the slope of the line segment AB, defined as the
ratio of the altitude over the base of the right triangle ABC. In
mathematics, slope is defined as rise over run; rise being the vertical
distance that has to be covered to move from point A to point B, and run
being the horizontal distance covered.
We have arrived at a very important conclusion: the average velocity of a
body as it makes a transition from A to B is the slope of the line segment
AB in the time-position graph. This conclusion is important because it
allows us to analyze the motion of a body using geometrical methods.
Since the slope in a time-position graph measures the average velocity, we
should look into the properties of the slope of a line segment.
The slope of a horizontal line is zero. So, if we find that the time-position
graph of a body is a horizontal line, we will immediately conclude that the
body is at rest, since zero slope indicates zero velocity. A vertical line, on
the other hand, suggests that the body moved from its initial position to
the final position in zero time, which produces an infinite slope. Since a
body cannot move with infinite velocity, we would not expect to see a
vertical time-position graph when we study the motion of a body. If the line
has an upward inclination to the right, its slope is positive, and we know
that the body is moving to the right. In a time-position graph with a
downward inclination to the right, the slope is negative, identifying a
velocity moving from right to left. These situations are displayed graphically
in the figure below.
Let us look into the geometrical interpretation of instantaneous velocity. We
have considered the two points, A(t1, x1) and B(t2, x2) on the time-position

Page 44 of 261

graph of the
motion of a
body,
computed the
average
velocity as the
body made the
transition from
A to B in the
time interval t
= t2 t1 and
established
that the
average
velocity v is the
slope of the
line AB. Let us
now consider
point C(t3, x3)
at a time t3, earlier than t2, as shown in the figure to the right. The average
velocity v during the transition from A to C in the time interval t = t3

t1 is the slope of the line AC. Since the line AC is more inclined towards the
vertical than is AB, the slope of AC has a larger value than does the slope
of AB. Thus, v > v.
In decreasing the time interval from
t to t we have moved the point B
to the point C on the time-position
graph. If we keep on decreasing the
time interval, point C will move on
towards point A along the graph, and
for each new value of t, we will get
a value of the average velocity that
will be different from the previous
one. We can see that if the timeposition graph of the motion of a
body is not a straight line, then the
average velocity of the body is not a constant. The value of the average
velocity will depend on the choice of t, the time interval.
What will happen if we bring point C as close to point A as possible without
making it coincide with point A? This action is equivalent to making t as
close to zero as possible without actually making it equal to zero. The line

Page 45 of 261

segment AC now becomes tangent to the time-position graph at point A(t1,


x1), and the slope of this tangent at A is the instantaneous velocity of the
body at time t1

The process of making a quantity, t in this case, as close to zero as


possible without making it equal to zero, is called a limiting process or
taking the limit as t approaches zero. We indicate this limiting process
by

and define the instantaneous velocity

v as

We interpret this formula as the value of

when t is as close to zero as

possible without being equal to zero. This limit has a finite value, equal to
the slope of the tangent to the time-position graph at point A(t1, x1).
The above discussion is a detailed explanation of the definition of
instantaneous velocity given by Equation 2-3 in the textbook.
Instantaneous velocity should be conceptualized as velocity v at a given
instant of time, distinct from average velocity v computed over a time
interval t.

Reading Assignment
Read Section 2-8 in the textbook.

Assignment for Credit


You should now complete Experiment 1 in the lab manual and send the lab
report to your tutor for marking. Getting feedback on the first report will
help you avoid repeating the same mistakes in the remaining experiments.

Assignment 1 Questions
7. A ball is released from top of 98 m tall building. At the same

Page 46 of 261

instant a second ball is thrown straight up from the ground with


an initial speed of 28 m/s directly below the point where the first
ball was dropped. Ignore air resistance.
a. At what height above the ground do the balls collide?
b. What are the balls velocities (magnitudes and
directions) just before the collision?
8. A moon rock is thrown upward with initial velocity 5.0 m/s. After
5.0 s it has a downward velocity of 3.0 m/s.
a. What is the acceleration due to gravity on the Moon?
b. How high above the starting point did the rock go
before it began to fall?
c. How high would the same rock rise above the earths
surface if it is thrown with the same speed?
Before starting the next lesson, test your comprehension of the material covered in
this unit by turning back to the objectives. Make certain that you can meet each
objective. If you have difficulties, do not hesitate to contact your tutor to discuss the
problem.

Footnote
1

Note that all references to the textbook are to Physics: Principles with
Applications, 6th ed., by Douglas C. Giancoli (Prentice Hall: Upper Saddle River, NJ:
2005).
Top

STUDY GUIDE

Lesson 5
Vectors and Scalars

n our discussions of displacement, velocity and acceleration, we briefly


mentioned vectors. A vector is a mathematical entity that can be defined
by two pieces of information: a numerical value (the magnitude of the

Page 47 of 261

vector) and a statement of the direction associated with the vector. In this
lesson, we consider some of the mathematical properties of vector
quantities.
If you think about it, you will realize that given the nature of the physical
quantities we deal with, such as displacement, velocity and acceleration,
someone had to invent the concept of vectors. Certainly, we need
numerical values to describe the motion of a body. However, numerical
values are not enough; the directions attached to these quantities have
crucial roles to play in their behaviour. A body speeds up when its velocity
and acceleration are in the same direction. When the two are in opposite
directions, the body slows down and eventually comes to rest. We,
therefore, need a mechanism that will allow us to deal with both of these
parametersthe numerical value and the directionsimultaneously.
Mathematicians came to the rescue with a new kind of mathematical object
designed to handle such situations, and named it a vector.
At this stage our interest centres around three vector quantities:
displacement, velocity and acceleration. As you get deeper into physics,
you will come across many other vector quantities. Because of their
importance in physics, it is essential that you learn how to handle vectors
before going any further in the course. In this lesson, you will learn the
basics of vector algebrahow to add vectors, how to subtract one vector
from another, and how to multiply a vector by a scalar.

Objectives
After completing this lesson, you should be able to
1. differentiate between a vector and a scalar quantity.
2. multiply a vector by a scalar quantity.
3. represent a vector graphically by a directed line segment.
4. add and subtract vectors using the tail-to-tip and parallelogram
methods.
5. decompose a vector into components.
6. find the resultant of multiple vectors using the components
method.

Introduction
We are all very familiar with numberswe cannot live without them. You

Page 48 of 261

need numbers to purchase your groceries, to tell the time and to determine
who will win the next municipal election. So, what is a number? For
example, what is a four? The number four is an idea in your mind. When
you were in school, your teacher showed you collections of four objects
four books, four toys, four people, etc. From these groups of four similar
objects you filtered out the idea of four-ness, gave it the name four, and
stored it in your mind. It is not a physical object, but you can attach
physical objects to it. To communicate about this idea, you give it a
graphical representation. You draw a picture (4), which is a symbol to
denote the idea of four-ness. In another country with a different language,
other people have given the same idea of four-ness a name of their own,
and they draw a different picture. For example, the symbol 8, which we use
for the idea of eight-ness is used by the Bengalees of Calcutta to represent
the number four.
The new mathematical objects (vectors) that you are going to learn about
are also ideas or abstractions. They become physical objects, such as
velocity or acceleration, when we attach appropriate units, such as m/s or
m/s2. To represent them graphically, we must simultaneously show the
magnitude and the direction of the vector quantity. We do so by drawing a
directed line segmenta straight line with an arrowhead at the end of it.
The length of the line segment equals the magnitude of the vector in some
chosen unit, and the orientation of the arrow shows the direction.
The representation of a vector by a directed line segment allows us to
develop the mathematical properties of vectors graphically in a simple but
elegant fashion. As you learn these properties, you will find that the
process of adding and subtracting vectors is markedly different from the
process of adding and subtracting numbers (or scalars), and this fact will
probably make you a bit more aware of the abstract nature of
mathematical operations.

Reading Assignment
Read Section 3-1 in the textbook, and make sure that you understand the
difference between vectors and scalars.1

Exercise
Give ten examples, from everyday life, of scalar quantities and another ten
examples of vector quantities.
Solution
Examples of scalar quantities include the distance to the moon, your bodys

Page 49 of 261

temperature, the speed of sound in air, the mass of an apple, the area of
your backyard, the volume of a cup of coffee, the calories in a donut, the
price of your new Plasma TV, the size of your computer hard disk and the
mark you get in your physics midterm.
Examples of vector quantities include the displacement of a golf ball after
the first shot, the wind velocity during a hurricane, the acceleration of a
rocket launched into space, the force applied to tow your car, the electric
field below a high-voltage power line, the Earths magnetic field at your
location, the force of gravity acting on a balloon, the momentum (to be
introduced later) of a truck just before a collision, the velocity of a
swimmer in a 100m freestyle competition, and the change in your position
when you take the elevator to the sixth floor.

Graphical Representation of a Vector

Top

The simplest way of representing a vector


a representation that will contain the
numerical value and the direction
simultaneouslyis to use a directed line
segment, as shown. The length of the line,
in a chosen unit, represents the magnitude
of the vector, and the arrowhead shows the
direction. The end with the arrowhead is called the tip of the directed line
segment, and the other end is called the tail. Bold letters (sometime with

an arrow on top), such as A, B, a, b, etc. (or


A,
B,
a,
b,
etc.), are
normally used to denote vector quantities, while italicized letters are used
for scalars. As the magnitude of a vector is a scalar quantity, it is common
to use A for the magnitude of the vector A.

Equality of Vectors
Two vectors A and B are said to be equal if they have equal magnitudes
(i.e. A = B and if both of them point in the same direction. Graphically, the
vectors A and B will be equal if the line segments representing the two
vectors are parallel and have the same length. Only then we can conclude
that the two vectors are equal, and write the equality as A = B. Note that
the two vectors in the diagram above have equal magnitudes; that is, A =
B However, the vectors themselves are not equal; that is, A B because
each has a different direction.
The velocity vector for a body moving at 60

Page 50 of 261

km/hr in the eastward direction, denoted


by VE, can be graphically represented by a
directed line segment of length six units,
each unit representing 10 km/h, with an
arrowhead pointing east. Similarly, the
vector VN, as shown in the diagram,
represents a velocity vector of magnitude
60 km/h in the northward direction; that is,
it has a length of six units and is directed to
the north. Although the magnitudes of the
two vectors are the same, that is VE =
VN, the vectors themselves are distinctly different, because they have
different directions.
This example brings up an interesting point. If we move the directed line
segment from one place to another without changing its direction, the
vector remains unchanged. Think about it.

The Negative of a Vector


The negative of a vector A, written as A, is the
vector whose magnitude is the same as that of
A, but whose direction is exactly opposite to that
of A. Graphically, the two directed line segments
representing A and A must have the same
length and must be parallel to each otherwith
the arrowheads pointing in opposite directions.
Now that we are familiar with the graphical representation of vectors, we
will see how such representations help us in mathematical manipulations of
vectors. A word of caution here: it is possible that in your prior studies you
were introduced to vectors in a different way. Some school textbooks have
a tendency to define vectors as directed line segments. This statement is
not accurate! A directed line segment is not a vector, but a representation
of a vector. Similarly, the notation $5 is not a five-dollar bill. You should
think of a vector as a physical entitysuch as displacement, velocity or
accelerationor as a mathematical object. We draw a directed line
segment to denote a vector quantity on a piece of paper.

Vector AdditionGraphical Method

Page 51 of 261

Top

A mathematical operation, such as addition, is a process of combining two


mathematical quantities to get a single mathematical quantity of the same
kind. In writing 3 + 5 = 8, we are making the following mathematical
statement: combining two numbers 3 and 5, through the process of
addition, we get the single number 8. The mathematical operation of
addition of vectors is stated the same way. If C is the vector we get by
adding the vectors A and B, then we describe the operation as A + B =
C. The vector C is called the sum or resultant of A and B. Although we
use the same notations and terminology as in addition of numbers, the
process of adding vectors is a bit more complicated. We do not just add the
magnitudes of two vectors to get their sum; instead, vector addition
requires a process that combines the magnitudes and directions of A and B
simultaneously, to tell us the magnitude and direction of C.

Vector AdditionTail-to-Tip Method


The tail-to-tip method defines the process of adding two (or more) vectors.
If two vectors are aligned from tail to tip to represent the two adjacent
sides of a triangle, then the third side gives the magnitude and direction of
the sum of the two vectors.
To see how it works, consider two vectors, A and B, as shown in the
diagram, below.

Let us move the vector B, keeping its length and direction unchanged, until
its tail coincides with the tip of A, as shown in the top half of the diagram.
We can now visualize the two line segments as the two adjacent sides of a
triangle. If we now draw a line segment from the tail of A to the tip of B,
we have a complete triangle whose adjacent sides are A and B. The third
side that we drew to complete the triangle is then the sum, C, of A and B.

Page 52 of 261

The length of the third side gives the magnitude of C, and the arrowhead,
as shown, gives the direction.
We could have placed the tail of A at the tip of B, and then completed the
third side of the triangle, as shown in the bottom half of the diagram. The
third side gives the magnitude and direction of the sum, C, of A and B.
Notice that the sum of two vectors is independent of the order in which
they are added. In short A + B = B + A = C
The process of adding vectors allows us to replace two vectors, A and B, by
a single vector, C, where C = A + B. This concept is quite useful in physics
in simplifying problems involving a number of vectors, as you will see later.
We can add more than two vectors by repeating the procedure above
several times. Consider the four vectors A, B, C and D, shown in the
diagram below.

To find the sum of these four vectors, we must repeat the tail-to-tip
alignment procedure a few times. First, we will find A + B by moving B
until its tail coincides with the tip of A; then, we complete the triangle by
drawing a directed line segment from the tail of A to the tip of B. Next, we
put the tail of C at the tip of the vector A + B, and complete the triangle to
find A + B + C. Finally, we add D to the sum A + B + C, placing the tail of
D at the tip of A + B + C, and completing the third side, which gives A + B
+ C + D.
In general, if n vectors can be represented by the adjacent sides of an open
(n+1)-sided polygon, then the last side gives the magnitude and direction
of the sum of the n vectors. As shown in the diagram above, the four
vectors, A, B, C and D, are the four adjacent sides of an open five-sided
polygon. The last side, the one that closes the polygon, gives the
magnitude and direction of the vector sum A + B + C + D. We would like
to stress that the vectors do not have to be added in any particular order.
As an exercise, use the polygon law to find D + B + A + C and B + C + A
+ D, in those orders, and see if you get the same sum (or resultant).

Vector AdditionParallelogram Method

Page 53 of 261

According to the tail-to-tip method, the third side, of the triangle whose
two adjacent sides are A and B is the sum of the vectors A and B. We have
mentioned that if we move a directed line segment from one place to
anotherwithout changing its length or directionwe still have the same
vector. So, if we move B from the tip of A to its tail, we have not changed
anything. The sum of A and B remains unchanged. If we complete the
parallelogram, with A and B as the two adjacent sides, we recognize that A
+ B is the diagonal of the parallelogram. We can say then that if two
vectors can be represented by the two adjacent sides of a parallelogram,
then the diagonal of the parallelogram gives the magnitude and direction of
the sum of the two vectors. This is known as the parallelogram method of
addition of vectors. It is obvious that the parallelogram method is a
variation of the tail-to-tip method of addition of vectors.

Subtraction of Vectors
Subtracting vector B from vector
A is equivalent to adding vector A
to the negative of vector B; that
is A B = A + (B). We can
then use the tail-to-tip or the
parallelogram methods to find A
+ (B). The diagram below
shows how one uses the tail-totip method to find A B for the
given vectors A and B. The difference between A + B and A B can also
be seen in the diagram. As an exercise, use the graphical method to find B
A and see how it compares with A B.

Reading Assignment
Read Sections 3-2 and 3-3 in the textbook.

Chapter 3: problem 2 (Page 65)


A delivery truck travels 18 blocks north, 10 blocks east, and 16 blocks
south. What is its final displacement from the origin? Assume the blocks
are of equal length.
Solution
We choose the north and east directions as shown in the diagram to the
right. To find the final displacement of the truck, we first represent each of

Page 54 of 261

the three displacements by an arrow with


appropriate length and direction and then
use the tail-to-tip method to arrange the
arrows as shown in the figure. The final
(or resultant) displacement R is
represented by the thick arrow, which
starts from the tail of the first
displacement and ends at the tip of the
last one.
After completing the right triangle below
the vector R, we use the Pythagorean
theorem to calculate the magnitude of the
resultant displacement as follows:

From the same triangle we can also write

which gives

So, the final displacement of the truck from the origin has a magnitude
equals to 10.2 blocks and a direction making an angle 11.3 north of east.
Note that the truck would make the same displacement if it travels 10 blocks east and
then 2 blocks north. Theseas we will see in the next sectionare called the
components of the vector R.

End of solution

Vector AdditionComponents Method

Top

Any number can be expressed as the sum of two numbers. For example,
the number 8 can be defined as the sum of 6 and 2. It is also the sum of 4
and 4. Actually, we can find an infinite number of pairs (3 and 5, 100 and
92, 1.4 and 6.6, and so on) whose sum will be 8.
We can do the same for vectorsbreak up a given vector into the sum of
two vectors. The process is called the decomposition of the vector. Let V
be a given vector represented by the directed line segment ac, as shown in

Page 55 of 261

the diagram below. If we draw the triangle abc we can say, following the
tail-to-tip method, that V = V1 + V2, where V1 and V2 are the vectors
represented by the directed line segments ab and bc, respectively. Since
unlimited number of triangles can be drawn, by changing the position of
point b only, we can find an infinite number of pairs of vectors whose sum
is V. In this decomposition, the vectors V1 and V2 are called the
components of vector V.
Components are extremely useful
tools in solving problems involving
vectors. In most cases, we choose
a pair of components that are
perpendicular to each other, from
the infinitely many choices
available. As we shall see,
perpendicular components make
life much easier when we must
deal with a multiple vectors.
Normally, it is more convenient to
choose horizontal and vertical
directions when we decompose a vector, unless the problem demands that
we choose other directions. The horizontal direction is labeled the x-axis,
and the vertical direction the y-axis. The components, Vx and Vy, along
these two directions are called the x-component and the y-component,
respectively. So, using the tail-to-tip method, we can also write V = Vx +
Vy.
Suppose V makes an angle with the x-axis. The magnitudes of V, Vx and
Vy are the lengths of the line segments ac, ad and dc, respectively. Since
these three line segments are the sides of a right triangle, we can use
Pythagorean theorem
(ac)2 = (ad)2 + (dc)2
to derive the relation

From the definitions of the trigonometric functions (see Appendix A-7 in the
textbook), we can also write
and

Page 56 of 261

leading to the following relations:


Vx = V cos

and Vy = V sin

In these equations, it is very important to note that is the angle between


the vector V and the positive direction of the x-axis. We can redo the
calculations above using a different anglefor example, the angle

between the vector V and the positive direction of the y-axis. In this case
we get the following relations between the x- and y-components of V:
Vx = V sin

and Vy = V cos

The simplicity of the equation between the magnitude of the vector and its
mutually perpendicular components is one of the reasons we choose to
work with such components.
Note also that we can write

Similarly, we also have

Before getting into an example to see how resolving vectors into


perpendicular components helps us in finding the resultant of two or more
vectors, let us straighten out a few notational conventions. We will use a
bold letter to denote a vector and an italic letter to denote its magnitude.
Thus Vx will denote the magnitude of the vector Vx, the x-component of the

vector V, A will denote the magnitude of the vector A, and so on. When
you write longhand, you should use an arrow on the top of a letter, i.e.
A
to denote a vector A, and a regular letter to denote its magnitude.

Chapter 3: problem 4 (Page 65)


If Vx = 6.80 units and Vy = 7.40 units, determine the magnitude and
direction of

V.

Solution

Page 57 of 261

Since both components of V are known, its magnitude is easily calculated


using the Pythagorean theorem such that,

To find the direction of V, it is convenient


to draw a diagram of the vector and its
components. From the diagram, we see
that V makes an angle with the x-axis.
Choosing the right triangle oac, we can
write

which gives the angle

The angle can also be calculated using the cosine or the sine of the angle,
such that

Note that, in the equation above, we used a value for V roundedas an intermediate
stepto four significant figures instead of three. The vector V, then, has a magnitude
equal to 10.0 units, and makes an angle 47.4 below the positive direction of the xaxis.

End of solution

Reading Assignment
Read Section 3-4 in the textbook.

Chapter 3: problem 10 (Page 66)

Top

Three vectors are shown in (textbook) Fig. 3-32. Their magnitudes are
given in arbitrary units. Determine the sum of the three vectors. Give the
resultant in terms of (a) components, (b) magnitude and angle with the x

Page 58 of 261

axis.
Solution
The first step in solving this problem
is to resolve each vector into its xand y-components. We will start with
vector A, which makes an angle of
28.0 with the positive direction of
the x-axis. From the diagram below,
we can see that the components of A
are given by:
Ax = 44.0 cos 28.0 = 38.85
Ay = 44.0 sin 28.0 = 20.66
Before calculating the components of vector B, we should note that it is
making an angle of 56.0 upwards from the negative direction of the xaxis. Therefore, the x-component is negative while the y-component has a
positive value, such that
Bx = 26.5 cos 56.0 = 14.82
By = 26.5 sin 56.0 = 21.97
Vector C has only a (negative) ycomponent, since it is pointing
directly downward. So, we write
Cx = 0
Cy = 31.0
If R is the resultant vector, such that
R = A + B + C then the components
of vectors A, B and C must add up
to give the corresponding
components of the resultant vector R, such that
Rx = Ax + Bx + Cx = 38.85 14.82 + 0 = 24.03
Ry = Ay + By + Cy = 20.66 + 21.97 31.0 = 11.63
a. Therefore, the components of the resultant vector, rounded to
three significant figures, are

Page 59 of 261

Rx = 24.0 and Ry = 11.6


b. The magnitude of R can be easily calculated using the
Pythagorean theorem:

From its vector representation, shown above, we see that R


makes an angle of

above the positive direction of the x-axis.


End of solution
Chapter 3: problem 11 (Page 66)
Determine the vector
A
32 (previous problem).

C,

given the vectors

Solution
Assume that the vector D is equal to the
difference between vectors A and C, from the
previous problem, such that D = A
C. Then, the components of D are equal to
Dx = Ax Cx = 38.85 0 = 38.85
Dy = Ay Cy = 20.66 (31.0) = 51.66
The magnitude of D is, then, equal to

making an angle

above the positive direction of the x-axis.


End of solution

Page 60 of 261

and

in (textbook) Fig. 3-

Chapter 3: problem 56 (Page 69)


What is the y component of a vector (in the xy plane) whose magnitude is
88.5 and whose x component is 75.4? What is the direction of this vector
(angle it makes with the x-axis)?
Solution
In this problem we are only given the
magnitude (88.5) and the x component
(75.4) of the vector. This information
allows for two possible orientations of the
vector, as shown in the diagram. Both V1
and V2 have the same magnitude and the
same value for x-component; the ycomponents have opposite signs.
Mathematically, we can see that two
possible solutions exist if we solve the
equation
V2 = Vx2 + Vy2
which relates the magnitude of a vector to its componentsfor Vy:

So, the y-component of V1 is 46.3, and the y-component of V2 is 46.3.


The angle , that each vector makes with the x-axis is equal to

Thus, the problem has two solutions as shown in the diagram above. The
first solution, V1, has a positive y-component and makes an angle of 31.6

x direction. The second solution, V2, has a negative y


component and makes an angle of 31.6 below the positive x direction.

above the positive

End of solution

Chapter 3: problem 68 (Page 71)

Page 61 of 261

a. A skier is accelerating down a 30.0 hill at 1.80 m/s2 (textbook


Fig. 3-48). What is the vertical component of her acceleration?
b. How long will it take her to reach the bottom of the hill, assuming
she starts from rest and accelerates uniformly, if the elevation
change is 335 m?
Solution
a. To compute the vertical (or y) component of the acceleration, we
first complete the vector diagram as shown below. If (a = 1.80
m/s2) is the magnitude of the skiers acceleration down the hill,
then using simple trigonometry, we see that the vertical
component of the acceleration is equal to

ay = a sin 30.0 = (1.80 m/s2) sin 30.0 = 0.900 m/s2


b. If the change in elevation is 335 m, then the distance covered by
the skier along the slope, from point A to point B, is equal to

Let t be the time


required by the skier
to cover the 670 m
along the slope
starting from rest and
moving with an
acceleration of 1.80 m/s2. Using the kinematic equation

and substituting the values


1.80 m/s2, we get

d = 670 m, v0 = 0, and a =

The skier will, therefore, take a time

to reach the bottom of the hill.

Page 62 of 261

End of solution

Exercise
Solve Problems 5, 8, 9 and 13 at the end of Chapter 3 in the textbook
(pages 65-66).
Answer to 8: (a) (6.6, 0) and (6.0, 6.0). (b) 6.0, 96 with the positive xaxis.

Assignment 1 Questions
9. A ship sails 130 km due north from island A to island B and then
92 km, in the direction 22 south of east, to island C. The ship
after that returns directly to island A. Calculate the magnitude
and direction of the displacement vector in the last trip. Draw
appropriate diagrams.
10. Three vectors, A, B and C each have a magnitude of 50 units.
Their directions relative to the positive direction of the x-axis are
20, 160 and 270, respectively. Calculate the magnitude and
direction of each of the following vectors.

a.
A + B + C

b.
A B + C

c. 2 (
A + C )

Before starting the next lesson, test your comprehension of the material covered in
this unit by turning back to the objectives. Make certain that you can meet each
objective. If you have difficulties, do not hesitate to contact your tutor to discuss the
problem.

Footnote
1Note

that all references to the textbook are to Physics: Principles with Applications,
6th ed., by Douglas C. Giancoli (Prentice Hall: Upper Saddle River, NJ: 2005).
Top

STUDY GUIDE

Page 63 of 261

Lesson 6
Projectile Motion

ave you closely watched the flight path of a baseball? It follows a


curved path, soaring up until it reaches its maximum height and then
descends gracefully, along a curved path, until it hits the ground. The
motion of the baseball is called projectile motion. Simply put, when a
body is projected into space, the subsequent motion is a projectile motion.
When a batter hits a baseball, the baseball is projected into space, and the
result is a projectile motion. If you throw a rock, shoot a gun or fire a
cannon, you are generating a projectile motion of some object. If the air
resistance is negligible, the path of a projectile in motion is always a
parabola or a part of a parabola.
It is natural to expect the analysis of motion along a curved path to be a bit
more complicated than the motion along a straight line as we considered in
previous lessons. However, as you will see in this lesson, projectile motion
can be analyzed in a simple and elegant way, requiring only our knowledge
of kinematics of linear motion and vector addition.
The procedure is as follows. First, we resolve the initial velocity into two
componentsalong the horizontal and vertical directions. Then, we analyze
motion of the body in each direction as linear motion independent of the
motion in the other direction. Since the acceleration due to gravity is
vertical (pointing downward), it will only affect the vertical component of
the velocity. The horizontal component, on the other hand, is not affected
and stays constant during the whole motion.

Objectives
After completing this lesson, you should be able to
1. describe the horizontal part of the motion of a projectile.
2. describe the vertical part of the motion of a projectile.
3. calculate the position and the velocity of a projectile at any time
while in flight.

Reading Assignment

Page 64 of 261

Read Section 3-5 in the textbook, and make sure that you understand the
kinematic differences between the vertical and the horizontal motions of a
projectile.1
We will analyze the motion of a body projected with an initial velocity v0
making an angle with the horizontal direction. To make our work as
simple as possible, we will consider an ideal atmosphere in which there is
no air resistance or wind turbulence of any kind. At this stage, you are
learning how to set up and solve simple model problems, and should not
worry about the fact that it is very difficult to prepare such an ideal
environment. You can be sure, however, that when NASA sends out a
spaceship, they do consider air resistance, wind velocity, variations in the
acceleration due to gravity and a host of other variables. Our task here is
much simpler!

We resolve the initial velocity v0 into horizontal and vertical components,


as shown in the diagram above. The magnitudes of the two components
are

vx0 = v0 cos

and

vy0 = v0 sin

respectively. We will look at each of these components separately. Since


there is no acceleration along the x-axis (ax = 0), the horizontal component

of the velocity will not change with time; that is, vx = vx0 at all times. The
horizontal position of an object after a time t will then be

Page 65 of 261

x = x0 + vx0 t
where x0 is the initial horizontal position of the object. The vertical
component of the velocity will be affected by the downward acceleration of
gravity, where ay = g. Therefore, the kinematics of the motion in the ydirection is described by the following equations:

vy = vy0 + ay t,

y = y0 + vy0 t + ay t2

vy2 = vy20 + 2ay(y y0)

Consider the projectile shown in the diagram above, which started at point
a with an initial velocity v0 making an angle with the horizontal. Taking
point a to be the origin of the coordinate system, the initial position of the
object is given by x0 = y0 = 0. As shown in the diagram, after a time tb the

body will be at point b. The vertical displacement yb of the body will have
the value

yb = v0 sin tb g tb2
while the horizontal displacement of the body after a time tb will be

xb = v0 cos tb
The horizontal component of the velocity will remain unchanged, as there is
no acceleration along the x-axis. The vertical component of the velocity at b
will have a smaller value because the acceleration due to gravity will
decelerate the body. The value of the two components of the velocity after
time tb will therefore be given by

vxb = vx0 = v0 cos


vyb = vy0 g tb = v0 sin g tb1
So, we notice that the vertical velocity of the projectile decreases as it goes
up towards the maximum height. We will thus conclude that after a time tb,
the velocity of the projectile has a smaller magnitude than the original
velocity, and the velocity vector is more inclined towards the horizontal.

At time tc, the body will reach its maximum height, at point c, where the
vertical component of the velocity will become zero. Setting vy = 0 and t =

Page 66 of 261

tc in the kinematic equation vy = vy0 + ay t such that


0 = v y0

g tc

we get

At this time, the projectile will have attained its maximum height.
Denoting the maximum height by h, and setting vy = 0 and
the equation vy2 = vy20 + 2ay(y y0) such that

y = yc = h in

0 = vy20 2gh
we get

At this point, when the projectile has reached its maximum height, the
body starts a reversal of its previous motion. The body now has the same
horizontal velocity, vx0, but its vertical component is zero.
Because of the downward acceleration of gravity, the body will now start
acquiring a downward (negative) velocity. Thus, at the point d, the
resultant velocity will have a direction below the horizontal axis. With the
passage of time, the downward component will increase in value, and the
resultant velocity will be directed more and more away from the horizontal.
After a time te = 2tc, the body will be at the altitude from which it started.
The horizontal distance traveled by the projectile in this time will be

R is called the range of the projectile. Note that in deriving the above

Page 67 of 261

result, we used the trigonometric identity: sin 2 = 2 sincos.


The motion of the projectile will be such that the velocity vector is always
tangent to the path, as shown. Also, note that the range of the projectile R,
as given by the formula above, will be greatest when sin2 has the
maximum value 1. This happens when 2 = 90, that is, when = 45.
Thus, for a given value of v0, a projectile will have a maximum range when
projected at an angle of 45.
Remember, we have assumed that there is no air resistance. In real life,
however, air resistance is present, and does have an effect on the projectile
motion. In our analysis we assume that the horizontal component of the
projectile velocity remains constant. Air resistance makes this assumption
untrue. So the horizontal velocity will also decrease with time, the result
being that the range will not be greatest when the projectile is projected at
an angle of 45. It can be shown, with some sophisticated calculations, that
the angle of projection must be somewhat less than 45 for maximum
range.

Reading Assignment
Read Section 3-6 in the textbook, and work carefully through the examples
in this section.
In summary, we start with a projectile with an initial velocity v0 at an angle
with the horizontal. We resolve the initial velocity into two components,
vx0 and vy0 and analyze the motion in each direction separately. The

horizontal component, vx0 remains constant because of the absence of any


acceleration in that direction. The vertical component,vy0, directed

upwards, decreases in value with increasing time, because of the downward


acceleration due to gravity. The resultant velocity of the body during this
upward flight becomes tilted more and more towards the horizontal,
becoming perfectly horizontal when the projectile reaches its maximum
height h. From that point on, the vertical component starts increasing in
value (moving in the downward direction) while the horizontal component
retains its constant value. The resultant velocity thus becomes tilted below
the horizontal direction. After twice the time it takes the body to reach its
maximum height, the projectile reaches the altitude from which it was
projected. Its velocity has a magnitude v0 at this point, but it now makes
an angle of below the positive direction of the x-axis.

Chapter 3: problem 18 (Page 66)

Page 68 of 261

Top

A diver running 1.8 m/s dives out horizontally from the edge of a vertical
cliff and 3.0 s later reaches the water below. How high was the cliff, and
how far from its base did the diver hit the water?
Solution
Let the height of the cliff be
h, and let d be the horizontal
distance from the base of the
cliff to the point where the
diver hits the water. The
problem says that the
horizontal velocity of the
diver, as he or she dives off
the cliff, is 1.8 m/s, and that
the diver reaches the water
3.0 s later.
Before we make our
calculations, let us first
summarize the information
we have about the motion of
the diver in both the horizontal and the vertical directions.
Horizontal Motion Vertical Motion

x0 = 0

y0 = h

x=d

y=0

vx0 = 1.8 m/s2

v y0 = 0

ax = 0

ay = 9.80 m/s2

t=3s

t=3s

To study the vertical motion of the diver, we substitute into the kinematic
equation y = y0 + vy0t + ayt2 such that
0=

h + 0 + (9.80 m/s2) (3.0 s)2

which gives the height of the cliff,

h = 44 m
Calculating the horizontal distance d, moved by the diver before hitting the
water, is straightforward, since there is no acceleration in this direction. So,
we write

Page 69 of 261

d = vx0 t = 1.8 m/s 3.0 s = 5.4 m


End of solution

Chapter 3: problem 22 (Page 66)


A football is kicked at ground level with a speed of 18.0 m/s at an angle of
35.0 to the horizontal. How much later does it hit the ground?
Solution
The time it takes the ball to fall back to the ground is twice the time it
takes to reach the maximum height. So, to solve this problem, we first
calculate the time for the upward part of the motion, and then multiply by
two.
Since the vertical component of the velocity at the top of the motion is
equal to zero, we can summarize the information we have about the
vertical motion of the ball during the first half of its trip as follows:

By substituting into the kinematic equation vy = vy0 + ayt, we get


0 = 18.0 m/s sin35.0 + (9.80 m/s2)

The time it takes the ball to reach the maximum height then, is equal to

Coming down requires the same amount of time, so the football hits the
ground
2t = 2 1.0535 s = 2.11 s
after being kicked. Note that we rounded to three significant figures in the
last step.
End of solution

Chapter 3 problem 24 (Page 67)

Page 70 of 261

An athlete executing a long jump leaves the ground at a 28.0 angle and
travels 7.80 m.
a. What was the takeoff speed?
b. If this speed were increased by just 5.0%, how much longer
would the jump be?
Solution
a. The range (i.e., the horizontal distance) covered by a projectile
with an initial velocity v0 projected at an angle with the
horizontal axis is

For the athlete executing the long jump, = 28.0 and R =


7.80 m, so we have the following value for the take-off speed

b. An increase of 5% in the take-off speed would make v0 =


10.08 m/s. Therefore, using the range formula, the horizontal
distance traveled by the athlete would be

Therefore, the jump would be (8.60 m 7.8 m = 0.80 m) longer


(i.e. a 10% increase).
End of solution

Chapter 3: problem 27 (Page 67)

Top

The pilot of an airplane traveling 180 km/h wants to drop supplies to flood
victims isolated on a patch of land 160 m below. The supplies should be
dropped how many seconds before the plane is directly overhead?
Solution
When released, the supplies will have the same velocity as that of the

Page 71 of 261

plane. In other words, the supplies will have the following initial velocity
components

Since there is no acceleration in the horizontal direction, the supplies will


keep moving with the same horizontal velocity as that of the plane. This
means that the plane will always stay directly over the package until it hits
the ground.
So, we need only to calculate the time it takes the supplies to move the
vertical distance from the level of the plane to the ground. We start with a
summary of the information we have about the vertical motion of the
supplies.

Next, we substitute the information above into the kinematic equation

y = y0 + vy0 t + ay t2 such that


0 = 160 m + 0 + (9.80 m/s2) t2
where

Page 72 of 261

Therefore, the supplies must be dropped 5.7 s before the plane is directly
above the target. We can also add that the supplies must be dropped when
the plane is

x = 180 km/h 5.71 s = 50 m/s 5.71 s = 286 m


from the target.
End of solution

Chapter 3: problem 30 (Page 67)


A projectile is fired with an initial speed of 65.2 m/s at an angle of 34.5
above the horizontal on a long, flat firing range. Determine
a. the maximum height reached by the projectile,
b. the total time in the air,
c. the total horizontal distance covered (that is, the range), and
d. the velocity of the projectile 1.50 s after firing.
Solution
Since we know the magnitude and direction of the initial velocity, it is a
good idea to start by calculating the horizontal and vertical components of
the initial velocity of the projectile. These are equal to

a. Before calculating the maximum height reached by the projectile,


we will summarize the information we have about its vertical
motion up to the highest point, as follows:

Page 73 of 261

By substituting into the kinematic equation vy2 = vy20 + 2ay(y


y0) such that
0 = (36.93 m/s)2 + 2 (9.80 m/s2)

we get the following value for the maximum height reached by


the projectile

b. By substituting the information above into the kinematic equation


vy = vy0 + ayt, such that
36.93 m/s = 0 + (9.80 m/s2)

we can calculate the time it takes the projectile to reach the


maximum height to be

So, the total time in the air is twice this value and is equal to
T = 2t = 2 3.77 s = 7.54 s
c. There are several ways to calculate the range of the projectile. A
simple way is to multiply the total time by the horizontal velocity
such that
R = vx0 T = 53.73 m/s 7.54 s = 405 m
d. The horizontal velocity of the projectile is constant
(vx = 53.73 m/s) during the whole trip. The vertical velocity after

t = 1.50 s is calculated from the equation vy = vy0 + ayt such that


vy = 36.93 m/s + (9.80 m/s2) 1.50 s = 22.23 m/s

Page 74 of 261

So, the velocity of the projectile 1.50 s after firing will have a
magnitude of

and (as seen in the diagram) makes an angle

above the horizontal.


End of solution

Chapter 3: problem 32 (Page 67)


A shotputter throws the shot with an initial speed of 15.5 m/s at a 34.0
angle to the horizontal. Calculate the horizontal distance traveled by the
shot if it leaves the athletes hand at a height of 2.20 m above the ground.

Solution
The horizontal distance traveled by the shot is calculated by multiplying the
constant horizontal velocity by the total travel time. So, initially, we need
to find the total time taken by the shot to rise to its maximum elevation
and then fall down to the ground. The information we have about the
vertical component of the motion is summarized as follows:

Using this information, the travel time can be calculated directly if we


substitute into the kinematic equation y = y0 + vy0 t + ay t2 such that

Page 75 of 261

0 = 2.20 m + 8.667 m/s

t + (9.80 m/s2) t2

Which is then simplified as follows:


4.90 t2 8.667

t 2.20 = 0

Solving the quadratic equation above for t, we get

Since negative time is not realistic, the total travel time of the shot is

t = 1.99 s
Another way to find t is to calculate the time required for the shot to rise to
its maximum height, and then the time it takes to fall back to the ground.
This method is slightly longer, but should give the same result.
We can now calculate the horizontal distance traveled by the shot as
follows:

d = vx0t = (v0 cos ) t = 15.5 m/s cos 34.0 1.99 s = 25.6 m


End of solution

Chapter 3: problem 34 (Page 67)

Top

Revisit Conceptual Example 3-7 (in the textbook), and assume that the boy
with the slingshot is below the boy in the tree (Fig. 3-36 in the textbook),
and so aims upward, directly at the boy in the tree. Show that the boy in
the tree makes the wrong move by letting go at the moment the water
balloon is shot.
Solution
In this problem, the boy sitting on the ground, at point O, aims a slingshot
at the other boy hanging from a tree branch at point A (see the diagram
above). At the same instant as a water balloon is fired from the slingshot,
the boy in the tree lets go, starting a free fall. Our goal is to show that both
the water balloon and the falling boy reach point B at the same time.
Since the horizontal velocity of the balloon is constant, the time it takes the
balloon to cross the vertical line AC (i.e., to move a horizontal distance d) is
given by

Page 76 of 261

where the trigonometric relation

has been used in the derivation.

The next step is to find the vertical positions of the water balloon and the
boy after this time. To do that, we substitute the information appropriate to
the motion of each object into the kinematic equation y = y0 + vy0t +

at2. For the water balloon, which is shot from ground level (y0 = 0), we
have

Since the boy, releases himself (vy0 = 0) from an initial height h, we can
write

However, from the first equation, above, we see that


vertical position of the boy at

t = td is given by

Page 77 of 261

h = v0 sin td. So, the

So, we conclude that the balloon and the boy will reach point B at the same
time.
End of solution

Chapter 3: problem 65 (Page 70)


Spymaster Paul, flying a constant 215 km/h horizontally in a low-flying
helicopter, wants to drop secret documents into his contacts open car,
which is traveling 155 km/h on a level highway 78.0 m below. At what
angle (to the horizontal) should the car be in his sights when the packet is
released (Fig. 3-46 in the textbook)?
Solution
First, let us calculate the time it takes the packet to reach ground level.
Since the packet was released (vy0 = 0) at an initial height 78.0 m above
the ground, we substitute into the kinematic equation
such that

y = y0 + vy0t + at2

0 = 78.0 m + 0 + (9.80 m/s2) t2


The packet then takes

to fall to the ground.

During this time, the car moves (with constant velocity

Page 78 of 261

vc = 155 km/h = 43.06 m/s) from point D to point E, traveling a distance


dc = vc t = 43.06 m/s 3.99 s = 171.8 m
Since the helicopter moves with the same horizontal velocity as the packet
(i.e. vh = 215 km/h = 59.72 m/s), it will always be located directly above
the packet during its flight, as we saw in Problem 27. So, when the packet
meets the car at point E, the plane will be located directly above the car (at
point C). The horizontal distance covered by the packet (or the plane)
during this time is equal to

dp = vp t = (59.72 m/s) (3.99 s) = 238.3 m


From the right triangle ABD (in the diagram) we see that the car must be
at an angle

below the horizontal when the packet is released.


End of solution

Exercise
Solve Problems 17, 19, 23, 31, 67 and 75 at the end of Chapter 3 in the
textbook (pages 66-71).

Assignment 1 Questions
11. A stone is projected at a cliff of height h with an initial speed of
40.0 m/s, directed 60.0 above the horizontal. The stone hits the
top of the cliff 6.10 s after it is launched. Calculate
a. the maximum altitude of the stone.
b. the height of the cliff.
c. the stone

Page 79 of 261

12. A boy kicks a ball horizontally near the edge of a boardwalk, with
an initial speed of 9.0 m/s. A blowing wind gives the ball a
constant horizontal acceleration of 12 m/s2. The ball falls into the
water directly under the boy. Ignore the effect of air resistance
on the vertical motion of the ball.
a. Determine the height of the boardwalk above water.
b. If the blowing wind reverses direction while maintaining
the same strength, where does the ball fall when it is
kicked with the same initial speed?
Before starting the next lesson, test your comprehension of the material covered in
this unit by turning back to the objectives. Make certain that you can meet each
objective. If you have difficulties, do not hesitate to contact your tutor to discuss the
problem.

Footnote
1Note

that all references to the textbook are to Physics: Principles with Applications,
6th ed., by Douglas C. Giancoli (Prentice Hall: Upper Saddle River, NJ: 2005).
Top

STUDY GUIDE

Lesson 7
Newtons Laws of Motion

n our studies of kinematics in the previous lessons, we avoided the


question, What causes motion? Why does a body move the way it does?
What makes a body, initially at rest, start to move? What causes a moving
body to decelerate and eventually come to a stop? We will begin to look for
answers to these questions in this lesson, our starting point in the study of
dynamics.
We owe our insight into this field to the great seventeenth-century
scientist, Sir Isaac Newton (1642-1727). Newton was born in the same

Page 80 of 261

year that another great scientist, Galileo, died. In the summer of 1665, the
Great (bubonic) Plague was devastating England, and the universities and
schools were closed to prevent its spread. Newton, then a 22-year-old
student at Trinity College of Cambridge University, returned to his mothers
farm in Lincolnshire, and spent the next 18 months studying mathematics,
mechanics and optics. So began one of the most productive periods of
achievement by a single individual in the history of science. The period
culminated about 22 years later, with the publication of Newtons major
work, Philosophiae Naturalis Principia Mathematica, in 1687. Most scientists
consider it to be the most important book ever written in the field of
physics. During this short time, Newton had invented differential and
integral calculus, developed a new theory of light, discovered the universal
law of gravitation, and put together the ideas of his three laws of motion.
These three laws, the subject matter of this lesson, still serve as the basis
of dynamics, some 300 years later.
In Newtons description of nature, force plays the crucial role in
determining the motion of a body. In an ordinary sense, we think of a force
as a push or a pull. That notion works quite well, but as physicists we
should be more precise. The three laws of Newton allow us to do just that
they provide a precise mathematical definition and understanding of force,
the cause of motion. In this lesson we will learn the definitions of the three
laws, known as Newtons First, Second and Third Laws of Motion. In the
following lessons, we will see applications of these laws in solving problems
of dynamics.

Objectives
After completing this lesson, you should be able to
1. define and use Newtons First Law of Motion.
2. define and use Newtons Second Law of Motion.
3. define the mass of a body, and explain the relationships among
the inertia, mass and weight of a body.
4. define and use Newtons Third Law of Motion.
5. distinguish between action and reaction forces.

Newtons First Law of Motion


From our experience, we know that if we leave a book on a tabletop and do
not disturb it, it will stay there forever. Imagine a spaceship moving with

Page 81 of 261

constant speed in the vacuum of the outer space and distant from any
celestial object. Can you think of any reason why it would ever slow down
and eventually come to a stop?
Newton probably thought along the same lines while formulating his First
Law, which states:
A body at rest will remain at rest till some external agent acts on
it to make it move. A body moving with a constant velocity along
a straight line will keep on moving with the same velocity until an
external agent acts on it to make it speed up, slow down or
change direction of motion.
The external agent that causes a change of state of the body is called a
force.
Newtons First Law of Motion states that in the absence of a force, a body
will continue in its state of rest or of uniform motionforever. What, then,
will happen when a force is applied to a body? Newtons Second Law takes
care of this question.

Reading Assignment
Read Sections 4-1 and 4-2 in the textbook.1

Newtons Second Law of Motion


If a body is at rest in the absence of a force, the application of a force will
make the body move. As we have seen in previous lessons, if a body
moves from its initial position of rest, it must experience acceleration.
Consider another situation. If a body, moving with a uniform velocity along
a straight line, is left undisturbed, it will continue to move with the same
velocity along a straight line. To bring the object to a stop, we need to
apply a force that resists the motion of the body, causing it to decelerate.
If the force is applied at an angle to the direction of motion, the body will
change its direction of motion.
Since the body will speed up, slow down or change direction only when it
experiences acceleration, we must conclude that the force is the cause of
the acceleration experienced by the body. We now have a causal
relationship: force causes acceleration. Force is the cause, and acceleration
is the effect.

Page 82 of 261

We can use common sense to figure out how this causal relationship works.
On a given body, a larger force would cause a larger acceleration than a
smaller force. In particular, if the force applied on the body is doubled, its
acceleration will be doubled, and if the force is tripled, the acceleration will
be tripled, and so on. The acceleration is then said to be proportional to the
applied force, a description that is represented mathematically as

aF
where a is the acceleration and F is the force. The symbol means is
proportional to. Now consider applying the same force on two bodies, one
heavier than the other. Once again, from common sense and experience,
we know that the heavier body would experience less acceleration than the
lighter one. Actually, with careful measurements, we will find that if the
mass is doubled, the acceleration is reduced to one-half, and it becomes
one-third if the mass is tripled, without changing the applied force.
Mathematically, this means

where m is the mass of the body. The relationship between a, F and


then be written as

m can

The unit of force is called Newton, abbreviated by N. It is defined such


that a force of 1 N acting on a body of mass 1 kg produces an acceleration
of 1 m/s2. In this definition, the constant that appears in the equation
above becomes equal to 1. After rearrangement, we have
F=

ma

This formula is the mathematical form of Newtons Second Law. This simple
relationship has remained the cornerstone of physics since Newton
discovered it some 300 years ago.

Reading Assignment
Read Sections 4-3 and 4-4 in the textbook, and work carefully through the
examples of the latter section.

Newtons Third Law of Motion

Page 83 of 261

Top

The one of Newtons laws most often quoted outside of physics is the Third
Law of Motion. We hear it everywhereaction and reaction are equal. A
physicist would quote Newtons Third Law as
To every action there is an equal and opposite reaction.
This statement appears simple, but do not be fooled! This law requires
quite a bit of thinking every time you want to use it. The point to remember
is that the action and reaction forces always act on different bodiesnever
on the same one. If you apply a force on the wall, the wall reacts with an
equal and opposite force on you. The wall experiences the action force, and
you experience the reaction force.
To see this point clearly, let us analyze what goes on when a book is left on
a tabletop, as shown in the diagram below. First, the force of gravity (or
weight)acts on the book in the downward direction; it is equal to mg. When
placed on the table, the book exerts a downward force F on the tabletop.
The tabletop, in return, exerts an upward reaction force FN (also called the
normal force) on the book. These two forces, F and FN, form the actionreaction pair. According to Newtons third law, the magnitudes of the two
forces are equal (i.e., F = FN), and the directions are exactly opposite.
Notice that the two forces of the action-reaction pair act on two different
bodies.

If we look at the book in isolation, we see two forces acting on it; its weight
mg and the normal force FN. Since the book is stationary and does not
move in the vertical direction, both forces must be equal in magnitude (i.e.,
FN = mg). Notice that both of these forces act on the book, and therefore
do not form the action-reaction pair. A better statement of the third law, to

Page 84 of 261

stress the point that action and reaction forces act on different bodies, is as
follows:
Whenever an object exerts a force on a second object, the
second object exerts an equal and opposite force on the first.
Newtons three laws form the basis of Newtonian mechanics. In the early
twentieth century it was discovered that Newtonian mechanics was not
exact for the motion of atomic and subatomic particles. A new theory,
called quantum mechanics was developed to explain the motion and
behaviour of such particles. The development of quantum mechanics,
however, does not invalidate Newtonian mechanics. For the macroscopic
world, where the objects of interest are large compared to the atomic and
subatomic objects of the microscopic world, Newtons laws form a
convenient and practical tool for understanding and describing the motion
of objects.

Reading Assignment
Read Sections 4-5 and 4-6 in the textbook.

Chapter 4: problem 2 (Page 98)


A net force of 265 N accelerates a bike and rider at 2.30 m/s2. What is the
combined mass of the bike and rider?
Solution
To solve this problem, we substitute the acceleration (a = 2.30 m/s2 of the
bike and rider into Newtons Second Law (F = ma). We know that the net
force is F = 265 N; therefore, we obtain
265 N =

m 2.30 m/s2

The mass of the bike and rider, then, is equal to

End of solution

Chapter 4: problem 4 (Page 98)


What is the weight of a 76-kg astronaut, traveling with constant velocity

Page 85 of 261

a. on Earth,
b. on the Moon (g = 1.7 m/s2),
c. on Mars (g = 3.7 m/s2), and
d. in outer space?
Solution
The mass of an object is a constant quantity that does not depend on the
location or any other external circumstances affecting the object. The
weight of an object, on the other hand, is not a constant quantity, because
it is equal to the mass times the gravitational acceleration affecting the
object in a particular location. Therefore, in this problem, the mass of the
astronaut is (m = 76 kg) anywhere, while his weight is as follows:
a. On Earth, where

g = 9.80 m/s2, we have

WEarth = 76 kg 9.80 m/s2 = 745 N


b. On the Moon, where

g = 1.7 m/s2, we have

WMoon = 76 kg 1.7 m/s2 = 130 N


c. On Mars, where

g = 3.7 m/s2, we have

WMars = 76 kg 3.7 m/s2 = 280 N


d. In outer space, where

g = 0, we have

WSpace = 76 kg 0 = 0
End of solution

Chapter 4: problem 8 (Page 98)

Top

A fisherman yanks a fish vertically out of the water with an acceleration of


2.5 m/s2, using very light fishing line that has a breaking strength of 22 N.
The fisherman loses the fish as the line snaps. What can you say about the
mass of the fish?
Solution
For the fishing line to pull the fish, the line must

Page 86 of 261

generate enough tension to counteract the


weight of the fish and to provide the upward
acceleration. The free-body-diagram on the right
shows the two forces acting on the fish while it is
being pulled. According to Newtons Second Law,
the net force acting on the fish is equal to the
mass times the upward acceleration, as follows:
T

mg = ma

and the mass of the fish is equal to

In this problem, the line snaps, which means that the tension required to
pull the fish is greater than 22 N. So, we conclude that the mass of the fish
is at least equal to

End of solution

Chapter 4: problem 15 (Page 98)


A person stands on a bathroom scale in a
motionless elevator. When the elevator
begins to move, the scale briefly reads only
75% of the persons regular weight.
Calculate the acceleration of the elevator,
and find the direction of acceleration.
Solution
The true weight of the person is always
equal to mg. However, the scale does not
measure the true weight directly, but rather
the contact (or normal) force between the
persons feet at the scales surface. From the
free-body diagram on the right, we see two
forces acting on the person: the persons
weight and the upward normal force. If the elevator is stationary or moving
with constant velocity, then the net force acting on the person must be
equal to zero, because there is no acceleration. In this case, the normal
force is equal to mg, and the scale measures the true weight of the person.

Page 87 of 261

In the case of upward acceleration, the normal force on the person


becomes greater than the true weight, to provide the necessary net force in
the upward direction. With downward acceleration, the normal force on the
person is smaller than the persons true weight.
In this problem, the apparent weight is only 75% of the persons true
weight. So, the normal force is equal to 0.75 mg, indicating that the
elevator is accelerating downward. From Newtons Second Law, we can
then write
Fnet =

mg 0.75 mg = ma

The mass cancels out, and the magnitude of the downward acceleration is
equal to

a = 0.25g = 2.5 m/s2


End of solution

Chapter 4: problem 18 (Page 99)


A person jumps from the roof of a house 3.9-m high. When he strikes the
ground below, he bends his knees so that his torso decelerates over an
approximate distance of 0.70 m. If the mass of his torso (excluding legs) is
42 kg, find
a. his velocity just before his feet strike the ground
b. the average force exerted on his torso by his legs during
deceleration.
Solution
a. The first part of this problem is a simple free fall problem with the
following information:

By substituting into the kinematic equation v2 = v02 + 2a(y y0),


we get

Page 88 of 261

v2 = 0 + 2 (9.8 m/s2) (0 3.9 m)


The final speed of the person just before his feet strike the
ground is, then, equal to

v = 8.7 m/s
b. When the persons feet touch the ground, the torso continues to
move down for a vertical distance of 0.70 m before it stops. The
information we have about the motion of the torso during its
deceleration is

By substituting into the kinematic


equation v2 = v02 + 2a(y y0), we get
0 = (8.743 m/s2)2 + 2
m)

a (0.70

So, while stopping, the torso


experiences an upward acceleration
equal to

a = 54.6 m/s2
At this stage, you may be tempted to multiply the acceleration
above by the mass of the torso to calculate the average force
exerted by the legs. However, you need to be careful here. From
the free-body diagram shown above it is clear that there are two
forces acting on the torso: its weight (mg) and the upward force
(Fleg) applied by the legs. Substituting the net force into
Newtons Second Law, we get
Fnet = Fleg

mg = ma

and

Page 89 of 261

End of solution

Exercise
Solve Problems 1, 3, 9, 12 and 14 at the end of Chapter 4 in the textbook
(pages 98-99).
Answer to 12: 3.8 m/s2 up.
Answer to 14: accelerate down at 2.2 m/s2.

Assignment 1 Questions
13. A 90 kg box is pushed by a horizontal force F at constant speed
up a ramp inclined at 28, as shown. Determine the magnitude of
the applied force.
a. when the ramp is frictionless.
b. when the coefficient of kinetic friction is 0.18.

14. Three boxes are in contact with each other on a frictionless


horizontal surface as shown. The masses of the boxes are
m1 = 10 kg, m2 = 20 kg, and m3 = 30 kg. A horizontal force

F = 90 N is applied to m1. Calculate.


a. the acceleration of the three boxes.
b. the net force on each box.
c. the contact forces between the boxes.

Before starting the next lesson, test your comprehension of the material covered in
this unit by turning back to the objectives. Make certain that you can meet each
objective. If you have difficulties, do not hesitate to contact your tutor to discuss the

Page 90 of 261

problem.

Footnote
1Note

that all references to the textbook are to Physics: Principles with Applications,
6th ed., by Douglas C. Giancoli (Prentice Hall: Upper Saddle River, NJ: 2005).
Top

STUDY GUIDE

Lesson 8
Applications of Newtons Laws

n this lesson, we look into how Newtons laws can be applied to


problems about the motion of a body. However, we do not want to give you
the impression that your study of applications of Newtons laws will end at
the end of this lessonwe will be using them throughout the course. Since
you have just learned these laws, it is appropriate to look at some simple
problems at this time, such as why it takes more effort to push a car uphill
than along a horizontal surface, how friction affects the motion of a body,
how a pulley system works, etc. We will use Newtons laws in more
complex situations as we progress in the course.
In applying Newtons second law, F = ma, we should note that the force F
is equal to the vector sum (or resultant) of all the forces acting on the
body. Some people are in the habit of writing the second law as Fnet = ma,

to remind themselves that one should take into consideration all of the
forces acting on the body when determining its acceleration. In any given
situation, therefore, we will take into account the gravitational force, the
frictional forcesif anyand the reaction forces, along with the applied
force, to compute the net force on the body.

Objectives
After completing this lesson, you should be able to

Page 91 of 261

1. distinguish between the force of static and the force of kinetic


friction.
2. draw a free-body diagram showing all the forces acting on a
particular object.
3. solve problems using Newtons laws of motion.

Reading Assignment
Read Section 4-7 in the textbook, and work carefully through the examples
in this section. We strongly recommend that you attempt the examples on
your own before looking at the solutions.1

Forces of Friction
Suppose you push a book placed on a horizontal tabletop. The book will
slide over some distance on the tabletop and then come to rest. Why
doesnt the book keep on sliding? When we push the book, we are exerting
a force. Once we stop pushing the book, the frictional force opposes the
motion and eventually brings the book to a stop. If there were no frictional
force, the book would keep on sliding with constant velocity, forever.
The term friction is commonly used for the force between two surfaces.
Experience tells us that moving a piece of furniture over a polished surface
is much easier than moving it over pavement. The reason, we say, is that
the friction between the furniture and the pavement is much larger than
the friction between the furniture and a polished floor.
We can think of the imperfections of the surfacesimperfections that resist
the motion of surfaces in contact with anotheras the cause of the
frictional force (see Fig. 4-26 in the textbook). When two surfaces are in
contact, the tiny protrusions of the imperfections push against each other,
opposing the motion of one surface over another. The sum of all these
forces can be thought of as the total frictional force. Although the model is
quite simple, it allows us to visualize why frictional forces between rough
surfaces are larger than those between comparatively smoother surfaces.
Let us imagine a wooden box on the floor. Suppose we apply a force on the
box in an effort to make it move forward. Assume that the applied force is
small, and the box will not move. This indicates that a force of friction
equal to and opposite to the applied force is generated when we apply the
force, making the net force on the box equal to zero. As the box does not
move, the frictional force thus generated is called the force of static
friction. As the applied force is increased, the force of static friction also

Page 92 of 261

increases, keeping the box in its place. Therefore, the static frictional force
is not a constant, but changes with the applied force. When the applied
force is sufficiently large, the box starts moving. We must conclude that
static friction increases with the applied force until it reaches a maximum
value, after which the box starts moving.
It has been verified experimentally that the maximum force of static friction
is proportional to the normal force between the surfaces in contact. This
relationship is translated mathematically into the following equation:
maximum Ffr = sFN
where FN is the normal force and Ffr represents the force of friction. The
constant of proportionality, s, is called the coefficient of static friction,
and its value depends on the nature of the surfaces in contact. Remember
that the equation above is true only for the maximum value of the static
frictional force between the two surfaces. If the applied force is smaller
than (sFN), then the force of static friction is equal to the applied force.
However, if the applied force is greater than (sFN), then there will be a net
force on the object, and the object will start to move. The correct
mathematical formula for static friction is then
Ffr sFN
with the equality sign holding when the body is on the verge of motion.
Once the box starts moving, the static friction vanishes and a new frictional
force, called kinetic friction, comes into the picture. You may have
noticed that it takes less effort to keep a body moving on a surface than to
start the body moving from rest. In other words, kinetic friction is less than
static friction. Kinetic friction also depends on the normal force and has the
mathematical form
Ffr = kFN
where k, the coefficient of kinetic friction, has a value that depends on the
surfaces and the velocity of the moving object.
Note: We provide this discussion to stress the point that frictional forces must not be
overlooked in computations of the net force acting on a body.

Reading Assignment

Page 93 of 261

Read Sections 4-8 and 4-9 in the textbook. Work out each of the examples
in Section 4-8, preferably attempting them on your own first.

Chapter 4: problem 22 (Page 99)

Top

A 650-N force acts in a northwesterly direction. In what direction must a


second 650-N force must be exerted so that the resultant of the two forces
points westward? Illustrate your answer with a vector diagram.
Solution
Let us represent the first force
by the vector A and the second
force by the vector B. Since the
first force acts in the
northwesterly direction, then
vector A makes an angle 45
up the negative direction of the
x-axis, as shown in the
diagram. Using the tail-to-tip
method, vector B is drawn such
that it starts from the tip of A
and ends somewhere on the xaxis. The resultant vector R,
which is directed along the
negative direction of the x-axis
from the tail of A to the tip of
B, has no y component. The line segments representing the three vectors
make an isosceles triangle, because A and B have equal magnitudes. So,
we conclude that is also equal to 45, and the second force must be
exerted in the southwesterly direction.
End of solution

Chapter 4: problem 26 (Page 100)


A person pushes a 14.0-kg lawn mower at constant speed with a force of
F = 88.0 N directed along the handle, which is at an angle of 45.0 to the
horizontal (Fig. 4-45 in the textbook).
a. Draw the free-body diagram showing all forces acting on the
mower.
b. Calculate the horizontal friction force on the mower.
c. Calculate the normal force exerted vertically upward on the

Page 94 of 261

mower by the ground.


d. What force must the person exert on the lawn mower to
accelerate it from rest to 1.5 m/s in 2.5 seconds, assuming the
same friction force?
Solution
a. See the free-body diagram on the right.
b. Since the lawn mower moves with constant
speed (no acceleration), the net horizontal
force acting on it must be equal to zero. So,
we write
F cos 45 Ffr = 0
The force of friction acting on the mower is
then equal to
Ffr = 88.0 N cos 45 = 62.2 N
c. The net force acting on the mower in the
vertical direction is also equal to zero, because there is no motion
in that direction. So we write
F sin 45 + FN

mg = 0

So, the normal force acting on the mower is equal to


FN = 88.0 N sin 45 + 14.0 kg 9.80 m/s2 = 199 N
d. For the mower to accelerate from rest (v0 = 0) to (v = 1.5 m/s)
in (t = 2.5 s), its acceleration must equal

Substituting the net horizontal force acting on the mower in


Newtons Second Law, we get
F cos 45 Ffr =

ma

where the applied force is an unknown in this part of the


problem. Using the values of Ffr and a from above, we get the
following value for the applied force:

Page 95 of 261

End of solution

Chapter 4: problem 34 (Page 101)


The two masses shown in Fig. 4-52 (in the textbook) are each initially 1.80
m above the ground, and the massless, frictionless pulley is 4.8 m above
the ground. What maximum height does the lighter object reach after the
system is released?
Solution
In this problem, the two masses are released at
the same height (1.80 m) above the ground. The
heavier mass (m2) accelerates downward and
the lighter mass (m1) accelerates upward with
the same acceleration. When m2 hits the ground,
the lighter mass will be at a height 3.60 m
above it. At this instant, the rope loses its
tension; however, m1 continues to move
upward, under the influence of gravity alone,
until it stops, momentarily, at the maximum height (h), as shown in the
diagram below.
To solve this problem, we start by studying the motion of the system
before the heavier mass hits the ground. Above are free-body diagrams
showing the forces acting on each mass. From Newtons Second Law, the
two equations describing the motions of the two masses are
T

m1g = m1a

and

m2g T = m2a
Since we are interested in calculating the acceleration of the system, we
may add the two equations above to get rid of T, obtaining
(m2

m1)g = (m1 + m2)a

The magnitude of the acceleration of each mass is equal to

Page 96 of 261

Now, we have enough information to calculate the velocity of the lighter


mass when the heavier one hits the ground:

By substituting into the kinematic equation v2 = v02 + 2a(y y0) such that

v2 = 0 + 2 1.815 m/s2 (3.60 m 1.80 m)


we get

v = 2.556 m/s
This is the velocity of both masses just before the heavier one hits the
ground. After that m1 will continue to move up, but under different
circumstances. It now undergoes a free fall with acceleration equal to that
of gravity. A summary of the information we have about the motion of m1
during this part of the motion is as follows:

Page 97 of 261

Substituting into the kinematic equation v2 = v02 + 2a(y y0) we get


0 = (2.556 m/s)2 + 2 (9.80 m/s2) (h 3.60 m)
The maximum height reached by the lighter mass is then equal to

End of solution

Chapter 4: problem 38 (Page 101)


Suppose that you are
standing on a train
accelerating at 0.20g. What
minimum coefficient of
static friction must exist
between your feet and the
floor if you are not to slide?
Solution
If the train is accelerating to the right, as shown in the diagram, then your
body must have the same acceleration to avoid sliding. This acceleration is
provided by the force of static friction between your feet and the floor of
the train. So, when the train is accelerating, the floor attempts to move to
the right under your feet. This means that you tend to slide to the left with
respect to the floor. Such a tendency, however, is resisted by the force of
static friction, which tries to keep you in the same spot on the trains floor.
If the coefficient of static friction is not large enough, then the acceleration
provided by the maximum force of static friction is less than the trains
acceleration, which causes you to slide backward. To avoid such a situation,
the maximum force of static friction must at least be equal to
Ffr = sFN =

ma

Since the normal force exerted by the floor on your body is FN =

Page 98 of 261

mg, we

then have
smg =

ma

Therefore, the minimum coefficient of static friction between your feet and
the floor must be equal to

Note that you did not need the numerical value of your mass to solve this problem.

End of solution

Chapter 4: problem 42 (Page 101)

Top

A car can decelerate at 4.80 m/s2 without skidding, when coming to rest
on a level road. What would its deceleration be if the road were inclined at
13 uphill? Assume the same static friction coefficient.
Solution

While moving on a level road, the net force acting on the car in the vertical
direction is equal to zero. So, as is clear from the free-body diagram above,
we have
FN

mg = 0

and, therefore, the normal force acting on the car is equal to its weight;

Page 99 of 261

that is,
FN =

mg

When the breaks are applied, the force of friction (Ffr) becomes the only
force acting on the car in the horizontal direction. In this situation, brakes
are applied just enough to allow the car to decelerate without skidding. In
other words, Ffr is equal to the maximum force of static friction. So, we
write
Ffr = sFN = smg
By substituting into Newtons Second Law, Ffr =
smg =

ma, we get

ma

The cars mass cancels out, and the coefficient of static friction is calculated
to be

Moving up an inclined road is a different scenario. First, the force of friction


is reduced because the normal force is now equal to mg cos , which is
smaller than the normal force on a level road. Second, there is the
additional force (mg sin ) caused by the component of the cars weight
down the incline. Therefore, the net force acting on the car, opposite to its
direction of motion, is equal to
Ffr +

mg sin = mg (s cos + sin )

By substituting into Newtons Second Law, we get

mg (s cos + sin ) = ma
The deceleration of the car up a 13-incline is then calculated to be

End of solution

Chapter 4: problem 48 (Page 102)

Page 100 of 261

Two crates, of mass 75 kg and 110 kg, are in contact and at rest on a
horizontal surface (Fig. 4-54 in the textbook). A 620-N force is exerted on
the 75-kg crate. The coefficient of kinetic friction is 0.15.
a. Calculate the acceleration of the system.
b. Calculate the force that each crate exerts on the other.
c. Repeat these calculations with the crates reversed.
Solution
a. To calculate the acceleration of
the system, we assume that
the two crates are glued
together and that they basically
form a single object of mass
(m1 +

m2) = 185 kg

Since the weight of the system


and the normal force (FN) are
the only two forces acting in
the vertical direction and since
they do not cause any
acceleration, we have
FN = (m1 +

m2)g

The force of friction acting to oppose the motion is then equal to


Ffr = kFN = k(m1 +

m2)g

and the net force acting on the system in the horizontal direction
is
620 N Ffr = 620 N k(m1 +

m2)g

By substituting into Newtons Second Law, for the whole system,


we get
620 N k(m1 +

m2)g = (m1 + m2)a

and the acceleration is calculated to be

Page 101 of 261

Since this is the final answer to this part of the problem, the
acceleration is rounded to two significant figures, and is written
as

a = 1.9 m/s2
b. From the diagram above, we see that the 620 N-force is applied
on the first crate. However, the second crate also moves and has
the same acceleration as the first crate. This means that the first
crate applies a force on the second crate and pushes it forward.
According to Newtons Third Law, however, the second crate
reacts by applying a force that has the same magnitude on the
first crate. These forces are called contact forces (FC) and can be
seen clearly if we draw free-body diagrams for the two crates
separately, as shown above.
To calculate FC, we can substitute the net force acting on the
second crate into Newtons Second Law, as follows
Fc

m2g = m2a

The force that each crate exerts on the other, is then equal to

As an exercise, see if you can get the same answer by applying


Newtons Second Law on the first crate.
c. This part is left for you as an exercise. The answer is Fc = 250 N.
End of solution
Top

Chapter 4: problem 64 (Page 103)


Suppose the coefficient of kinetic friction between

Page 102 of 261

m1 and the plane (

Fig. 4-57 in the textbook) is k = 0.15, and that

m1 = m2 = 2.7 kg.

m2 moves down, determine the magnitude of the acceleration


of m1 and m2, given = 25.

a. As

b. What is the smallest value of k that will keep this system from
accelerating?
Solution
a. We start by analyzing the forces acting on each mass, as shown
in the free-body diagram below. The first mass (m1) does not
accelerate in the direction perpendicular to the incline. So, the
net force acting on m1 in this direction is equal to zero, and we
can write
FN =

m1g cos

On the other hand, the mass

m1 experiences three forces acting

on it in the direction parallel to the incline: the tension in the


string pulling the mass up the incline, the component of the force
of gravity acting down the incline and the force of kinetic friction

Page 103 of 261

acting down the incline to resist the motion. By substituting the


net force into Newtons Second Law, we get the following
equation that describes the motion of m1 up the incline
T

m1g sin km1g cos = m1a.

The second mass (m2) has two forces acting on it in the vertical
direction: the weight m2g pulling down, and the tension in the
string trying to pull the mass up. Therefore, the motion of

m2 is

described by substituting the net force into Newtons Second Law


as follows

m2g T = m2a
Since we are interested in calculating the acceleration of the
system, we can proceed by adding the two equations above, thus
eliminating T, as follows

m2g m1g sin km1g cos = (m1 + m2)a


The acceleration of the system, is then equal to

b. To solve the second part of the problem, we set a = 0 and keep


k as an unknown to be determined. Then, we proceed as we did
in the first part, which results into the equation

m2 m1 sin km1 cos = 0.


Therefore, the smallest value of k that will keep the system from
accelerating is

Page 104 of 261

End of solution

Exercise
Solve Problems 23, 25, 32, 37, 45, 46, 53, 56 and 76 at the end of Chapter
4 in the textbook (pages 99-104).
Answer to 32: 25
Answer to 46: 6.7 kg
Answer to 56: 1.1 kg
Answer to 76: (a) 11.3 kg (b) 0.88 m/s2

Assignment 1 Questions
15. Three objects are connected on an inclined table as shown in the
diagram below. The objects have masses 8.0 kg, 4.0 kg and
2.0 kg as shown, and the pulleys are frictionless. The tabletop is
rough, with a coefficient of kinetic friction of 0.47, and makes an
angle of 20.0 with the horizontal.
a. Draw a free-body diagram for each object.
b. Determine the acceleration and direction of motion of
the system.
c. Determine the tensions in the two cords.

16. A 4.0 kg toboggan rests on a frictionless icy surface, and a 2.0 kg


block rests on top of the toboggan. The coefficient of static
friction s between the block and the surface of the toboggan is
0.58, whereas the kinetic friction coefficient is 0.48. The block is
pulled by a horizontal force of 30 N as shown.
a. Calculate the blocks acceleration.

Page 105 of 261

b. Calculate the toboggans acceleration


c. If the applied force is gradually reduced, at what value
do the block and the toboggan have the same
acceleration?

Before starting the next lesson, test your comprehension of the material covered in
this unit by turning back to the objectives. Make certain that you can meet each
objective. If you have difficulties, do not hesitate to contact your tutor to discuss the
problem.

Footnote
1

Note that all references to the textbook are to Physics: Principles with
Applications, 6th ed., by Douglas C. Giancoli (Prentice Hall: Upper Saddle River, NJ:
2005).
Top

STUDY GUIDE

Lesson 9
Uniform Circular Motion

ou must have done it when you were youngertied a rock or some


other object to the end of a string and rotated it around your finger. You
may even be doing it now, rotating a key chain or something around your
forefinger as you read. When you do this, you feel a pull on your finger.
The faster you rotate the piece of rock, the stronger the pull. If you let go
of the string, the object flies awayalong a straight line, forgetting the
circular path along which it had been rotating.
In this lesson, we explore aspects of uniform circular motion, including

Page 106 of 261

general characteristics, acceleration, force and the physics of curves.

Objectives
After completing this lesson, you should be able to
1. describe the motion of a body in a circular path.
2. define centripetal acceleration and centripetal force, and use
them to solve problems related to motion of an object along a
circular path.
3. distinguish centrifugal acceleration from centripetal acceleration.
4. solve problems involving banked and unbanked curves.

Kinematics and Dynamics of Circular Motion


To see what goes on when a body moves
along a circular path, let us look into the
circular motion of a body in a horizontal
plane, as shown in the diagram at the
right. We have chosen a horizontal plane
for simplicity. If the motion is confined to
a horizontal plane, the acceleration due to
gravity will be perpendicular to the
direction of motion, and will have no effect
on the motion of the body. Assume that
the magnitude of the bodys velocity at
point A is v. The direction of the velocity is
tangent to the circular path, as shown in
the diagram. According to Newtons First Law of motion, in the absence of
any force, the body will tend to move in the direction of the tangent to the
circle at A. To make the body move in a circular path and reach point B on
the circle, a force will have to be applied in a direction different from the
direction of the bodys velocity.
In general, we know that the velocity of an object is a vector quantity
which consists of a magnitude and a direction. A change in either the speed
or the direction of motion of the object causes a change in the velocity, and
we say that the object accelerates. According to Newtons second law, a
force is required to generate acceleration. If the force is applied in the
direction of the velocity, then only the magnitude will change, without
affecting the direction of motion. On the other hand, if the applied force is

Page 107 of 261

perpendicular to the direction of the velocity, only the direction of the


velocity will change, without changing its magnitude (or speed). In all other
situations, where the applied force is at an angle with the velocity, both the
magnitude and the direction of motion will change.
In a uniform circular motion, an object rotates with constant speed around
a circular path. So, the applied force must be perpendicular to the velocity
of the rotating object at all times. In other words, at every point of the
circular path of the body, the applied force must be perpendicular to the
tangent of the circular path pointing towards the centre. This is the force
that you apply through the string on an object when you make it rotate in a
circle. Since the applied force is directed towards the centre, it is called
centripetal force. The magnitude of the centripetal force is equal to

where m is the mass of the rotating object and r is the radius of the circular
path. The acceleration of the rotating object also points towards the centre.
It is called the centripetal acceleration and has a magnitude equal to

You may have heard of the concept of centrifugal force, which


supposedly pushes outward from the centre of an object in circular motion.
There is a widespread belief that the centrifugal force makes the object fly
away when you let go of the string. This belief is not true. The object at the
end of the string never experiences centrifugal force. So what does happen
when you let go of the string? The moment you let go, the source of the
centripetal force becomes non-existent. Since no force is acting on it
anymore, the object moves in the direction of v, according to Newtons first
law. Since the direction of v is always tangent to the circular path, the rock
flies in a direction tangent to the circle at the point of releaseit does not
continue along the radius.
The term centripetal force can be applied to any type of force that causes
circular motion. What we learn from the simple example of an object tied at
the end of a string also explains the movement of the planets around the
sun, the Bohr model of the atom and the motion of electrons and protons in
cyclotronsto give a few examples. Gravitational force provides the
centripetal force necessary for the motion of a planet in a closed orbit
around the Sun; the Coulomb force of electrostatic attraction is responsible
for the centripetal force that makes electrons orbit the nucleus in the Bohr
atom; and electromagnetic forces furnish the centripetal force that makes
the electrons and protons move in circular orbits in particle accelerators,

Page 108 of 261

with speeds close to the speed of light.


In future courses in physics, you will learn about the Bohr atom and about
the principles that make a cyclotron work. In the next lesson we look at
circular motion as it relates to gravitation.

Reading Assignment
Read Sections 5-1 and 5-2 in the textbook. Make sure that you understand
the meaning of a change in the velocity when the speed is constant and the
meaning of centripetal acceleration.1

Banked and Unbanked Curves

Top

Probably you have noticed that when you are driving on the highway,
within the speed limit and under normal driving conditions, that you are
able to negotiate most of the turns without the need to reduce your speed.
This is because civil engineers perform careful calculations while designing
each curve of the road so that vehicles will experience the centripetal force
needed to allow them to turn the curve while staying on the road.

Reading Assignment
Read Section 5-3 in the textbook.

Chapter 5: problem 1 (Page 130)


A child sitting 1.10 m from the centre of a merry-go-round moves with a
speed of 1.25 m/s. Calculate the centripetal acceleration of the child and
the net horizontal force exerted on the child (mass = 25.0 kg).
Solution
a. In this problem, the child performs a
uniform circular motion by moving in a horizontal circle of radius
r = 1.10 m, with constant speed equal to v = 1.25 m/s. The
centripetal acceleration of the child is, therefore, given by

b. As is clear from the free-body diagram to the right, there are

Page 109 of 261

three forces acting on the child. Only one of


these forces, the centripetal force, acts in
the horizontal direction. Therefore, the net
horizontal force has a magnitude equal to
FR =

maR = 25.0 kg 1.42 m/s2 =

35.5 N
and is always directed towards the centre of
rotation.
End of solution

Chapter 5: problem 6 (Page 130)


What is the magnitude of the acceleration of a speck of clay on the edge of
a potters wheel turning at 45 rpm (revolutions per minute) if the wheels
diameter is 32 cm?
Solution
To calculate the centripetal acceleration of the speck of clay, we must first
find its rotation speed. The information we have tells us that it rotates in a
circle of radius r = 0.16 m and makes 45 revolutions every minute. In
other words, during its uniform rotation, the speck of clay covers a distance
x = 45 2r in a time interval t = 60s. Therefore, the speed of the
speck of clay is

and its centripetal acceleration of is given by

End of solution

Chapter 5: problem 10 (Page 130)


How large must the coefficient of static friction be between the tires and
the road if a car is to round a level curve of radius 85 m at a speed of
95 km/h?
According to Newtons first law, the car has the
tendency to keep moving in a straight line, and therefore to go off track

Page 110 of 261

when the road bends. If the driver turns the


steering wheel slightly in the direction of the
curve, the tires will have to slide on the road to
allow the car to continue moving in a straight
line. This is exactly what we feel happening on
an icy road when there is little friction. However,
in normal driving conditions, the force of friction
between the tires and the road resists the sliding
motion, and the tires are only allowed to roll
forward along the road. Any motion
perpendicular to this direction is resisted by the
force of static friction, which also provides the
needed centripetal force that allows the car to turn while staying in the
same lane. Note that if the speed is too great, the required centripetal
acceleration may be larger than the maximum force of static friction can
provide. If so, the car will slide off the road.
The car shown in the free-body diagram above is moving into the page,
and is negotiating a level curve of radius r = 85 m bending to the left. For
the car to stay on the road, a centripetal force equal to

must act on the car towards the centre of the curve. The necessary force is
provided by the force of static friction, which is perpendicular to the
direction of motion of the car. It acts on the tires to the left, as shown in
the diagram. For the car to round the curve safely at a speed v = 95 km/h
= 26.39 m/s, the maximum force of static friction should be at least equal
to the required centripetal force; that is,

where in this case FN = mg. The coefficient of static friction must then have
a minimum value equal to

End of solution

Chapter 5: problem 14 (Page 130)


A sports car of mass 950 kg (including the driver) crosses the rounded top

Page 111 of 261

of a hill (radius = 95 m) at 22 m/s. Determine


a. the normal force exerted by the road on the car,
b. the normal force exerted by the car on the 72-kg driver, and
c. the car speed at which the normal force on the driver equals
zero.
Solution
While crossing a rounded top of a hill, a car performs uniform circular
motion in the vertical direction, as shown in the diagram below. Assume
that the car is at the top of the hill and moves forward with velocity v.
According to Newtons first law, the car tends to keep moving in a straight
line, and thus tends to lose contact with the road. The force of gravity acts
on the car in the downward direction and pulls the car down. The normal
force, on the other hand, acts on the car in the upward direction to prevent
it from penetrating the road. As a result, these two forces work together
and provide the required centripetal force to allow the car to cross the
rounded hill while staying on the road.

a. Since the centripetal force acting on the car while crossing the hill
is equal to the net force in the vertical direction, we can write

So, the normal force exerted by the road on the car is equal to

Page 112 of 261

b. To calculate the normal force on the driver, we follow a similar


procedure to that used above, because the driver performs the
same kind of motion as the car. The result is

Note that FN corresponds to the apparent weight of the driver.


Comparing this value to his actual weight, the driver feels that
he loses 52% of his weight while crossing the hill at this speed!

c. If the car crosses the same hill at a higher speed, the normal
force acting on the driver becomes smaller. At the right speed
there will be no normal force acting either on the driver or on the
car, and the driver will feel weightless. This happens when the
following equation is satisfied:

The speed of the car in this case is equal to

or approximately 110 km/h. Note that if the car exceeds this


speed, the force of gravity will not be able to provide enough
centripetal force to keep the car on the road.
End of solution

Chapter 5: problem 19 (Page 131)


A flat puck (mass M) rotates in a circle on a frictionless air-hockey
tabletop, and is held in this orbit by a light cord connected to a dangling

Page 113 of 261

Top

block (mass m) through a central hole as shown in Fig. 5-36 (in the
textbook). Show that the speed of the puck is given by

Solution
We start by drawing free-body diagrams for both masses, as shown in the
diagram below. Since the dangling block does not move, the net force
acting on it is equal to zero, or T = mg. The rotating puck performs uniform
circular motion, where the tension in the cord acts as the centripetal force.
So, we write

Solving the equation above for

v we get

End of solution

Chapter 5: problem 22 (Page 131)


A 1200-kg car rounds a curve of radius 67 m banked at an angle of 12. If
the car is traveling at 95 km/h, will a friction force be required? If so, how
much and in what direction?

Page 114 of 261

Solution
The free-body
diagram above
shows the car
moving into
the page and
turning left on
the banked
curve.

same lane), then a centripetal force equal to

If the car is to
turn at a speed
v = 95 km/h =
26.39 m/s
without
slipping (i.e.,
staying in the

is required, pointing to the left in the horizontal direction. If no friction is


needed, then the horizontal component of the normal force (FN sin )

provides the required centripetal force. In such case, the vertical


component of the normal force is balanced by the cars weight (i.e. FN cos
=

mg). As a result, the normal force acting on the car is given by

and the centripetal force provided is equal to

This value is much smaller than the centripetal acceleration required for the
car to round the curve without slipping. As a result, the car will tend to
move to the right, slipping up the banked curve. To keep the car on track,
static friction is required to act on the tires, down the plane, to resist the
slipping motion and supplement the needed centripetal force.

Page 115 of 261

If, after including the force of friction, the car is able to round the curve at
the indicated speed safely without slipping, then we can write the following
equations
FN cos Ffr sin

mg = 0

and

which correspond to the net forces acting on the car in the vertical and
horizontal directions, respectively. Since we are interested in finding Ffr, we
may substitute FN from the first equation above into the second equation
as follows:

Solving this equation for Ffr, we get

where we have made use of the trigonometric identity cos2 + sin2 = 1.


The quantitative value of the force of static friction down the plane is then
equal to

End of solution

Exercise
Solve Problems 3, 7, 9, 13, 16 and 21 at the end of Chapter 5 in the
textbook (pages 130-131).
Answer to 16: ( a) 1.7 m/s (b) 3.3 m/s

Assignment 1 Questions

Page 116 of 261

Note: These are the last questions of Assignment 1. After completing this assignment
you should submit your solutions to your tutor for marking by uploading it to the
appropriate drop box on the course home page. Given the requirements of the
assignments many students prefer to prepare handwritten solutions. In that case,
please submit a scanned copy of your solutions or a clear photograph. You also have
the option to send the assignment to your tutor via regular mail. Keep a copy of your
assignment, at least a rough draft, in case the original goes astray.
Although you may find it convenient to answer the assigned problems in the briefest
possible way, you are encouraged to get into the habit of showing all of your work. This
strategy enables the marker to identify where you are having trouble with concepts or
mathematical skills.
Also, note that the midterm exam material extends up to Lesson 11. Consult your
Course Manual for information about the examination. Your corrected assignment
should help you prepare.

17. A 0.60 kg sphere rotates around a vertical shaft supported by two


strings, as shown. If the tension in the upper string is 18 N.
Calculate
a. the tension in the lower string?
b. the rotation rate (in rev/min) of the system?

18. Earth is a spherical object, which completes a full rotation every


24 hours. How long should the day on Earth be so that an object
at the equator is able to float freely above the ground?
Before starting the next lesson, test your comprehension of the material covered in this
unit by turning back to the objectives. Make certain that you can meet each objective. If
you have difficulties, do not hesitate to contact your tutor to discuss the problem.

Page 117 of 261

Footnote
1

Note that all references to the textbook are to Physics: Principles with
Applications, 6th ed., by Douglas C. Giancoli (Prentice Hall: Upper Saddle River, NJ:
2005).
Top

STUDY GUIDE

Lesson 10
Gravitation

ntil the middle of the sixteenth century, theories of nature and of


planetary motion were dominated by the views of Aristotle and Ptolemy.
Most people believed that the Earth was the center of all planetary motion,
with the heavenly bodiesplanets, sun, stars, etc.moving in perfect,
circular orbits around it.
In 1543, Nicholas Copernicus (1473-1543), a Polish astronomer, published
the radical suggestion that the planets orbit the sunnot the earth. Some
50 years after the publication of Copernicus paper, a Danish nobleman,
Tycho Brahe (1546-1601), started making accurate observations of the
positions of the planets. After Brahes death, his assistant, the brilliant
German astronomer Johannes Kepler (1571-1630), started analyzing the
data and found, to his surprise, that they did not fit perfectly circular orbits
In 1609, Kepler proposed that planetary orbits are elliptical in shape.
A young astronomer, Edmund Halley (1656-1742), whose name is
associated with the famous comet, paid a visit to Isaac Newton in 1684.
Halley and a few colleagues from the Royal Society of London had been
working on the orbits of the planets around the sun. They had guessed,
from Keplers description of elliptical orbits, that the forces acting on the
planets must decrease with the square of their distances from the sun. But
Halley and his colleagues were unable to prove their conjecture. During his
conversation with Newton, then a professor at Cambridge University, Halley
inquired if Newton knew what the shape of an orbit would be if the force
were inversely proportional to the square of the distance. Newton replied
that under such forces, the orbits would be ellipses. On being asked by a
surprised Halley how he knew, Newton answered that he had calculated the

Page 118 of 261

solution.
Newton had reasoned that a planet follows a closed orbit because some
force makes it fall from the straight-line path it would follow in absence of
the forceand that successive falls from intended straight-line paths make
the resultant path a closed orbit. And, so the story goes, when he observed
an apple fall, he wondered if the force that made the apple fall was also
responsible for the fall of the planets from their intended straight-line
paths. These thoughts eventually came together and resulted in Newtons
famous Law of Universal Gravitation.
Newtons discovery of the Law of Universal Gravitation provides an
interesting insight into the world of scientific activity. Observations,
theories and ideas from various scientists and philosophers float around for
centuries as pieces of a puzzle, until a scientist arrives at the right time in
history, puts the various pieces together and makes the picture complete
a great theory is born. Of course, Newton did not suddenly stumble on the
law of gravitation! Ideas of attractive forces between different bodies had
been there from the early days of scientific reasoning, and several
scientists, mathematicians and philosophers, before and during Newtons
time, had ideas that could have led to the law of gravitation.
Newtons Law of Universal Gravitation, which is considered the crowning
achievement of his work, states that
Every object in this universe attracts every other object with a
force, along the line joining the two objects, that is proportional
to the product of the two masses and inversely proportional to
the square of the distance between them.
If two masses

m1 and m2 are separated by a distance r, then each one

exerts an attractive force F on the other along the line joining them. The
magnitude of the force is given by the equation

where G is known as the universal constant of gravitation. Its value has


been determined to be 6.67 1011 N m2/kg2

Objectives
After completing this lesson, you should be able to

Page 119 of 261

1. write
down the

mathematical equation describing Newtons Law of Universal


Gravitation.
2. use the Law of Gravitation and centripetal acceleration to solve
problems related to planetary motion and satellites.

Reading Assignment
Read Sections 5-6 and 5-7 in the textbook.1

Satellites

Top

A soldier was stationed on the top of a high mountain to guard new and
powerful artillery that was placed there for testing purposes. One day the
soldier decided to do a test himself. He positioned the artillery, pointing its
nozzle in the horizontal direction, and started firing shells into the vast
plans in front of him.
Since he was able to set the firing speed, he started with a relatively small
initial velocity and gradually increased it. When the firing speed was small,
relative to the size of the Earth, an observer would see something similar to
that in part (a) of the figure above. However, as the firing speed was
increased, an observer evaluating the test would have to take the
curvature of the Earth into consideration, as shown in part (b). If the
soldier kept increasing the firing speed, the shells would travel greater
distances around the globe before hitting the ground. Eventually, when a
shell was fired with the right speed, it would travel in a perfect circle
around the Earth without a decrease in its altitude. If this happened, the
soldier would have created a satellite that, if we neglect air resistance,
would keep rotating around the Earth, as shown in part (c). The soldier had
to be careful here, because if he did not change his location at the right
time, the shell would have hit his artillery from behind and destroyed it!

Page 120 of 261

Reading
Assignment
Read Section
5-8 in the
textbook,
paying special
attention to the
concept of
weightlessness

Chapter 5:
problem 28
(Page 131)
Calculate the
force of Earths
gravity on a
spacecraft
12,800 km (2
Earth radii)
above the
Earths surface
if its mass is
1350 kg.
Solution
The force of
gravity on the
spacecraft is
given by
Newtons Law
of Universal
Gravitation

In this problem, the different parameters in the equation above have the
following values:

Page 121 of 261

where r is the distance from the satellite to the centre of the Earth. By
substitution, we calculate the force of Earths gravity on the spacecraft as
follows:

End of solution

Chapter 5: problem 29 (Page 131)


At the surface of a certain planet, the gravitational acceleration g has a
magnitude of 12.0 m/s2. A 21.0-kg brass ball is transported to this planet.
What is
a. the mass of the brass ball on the Earth and on the planet?
b. the weight of the brass ball on the Earth and on the planet?
Solution
a. The mass of an object is a constant quantity that does not
depend on the location or the forces acting on the object. So, the
mass of the brass ball is equal to 21.0 kg on the Earth and on the
planet.
b. Since the weight is equal to the force of gravity acting on the
object, it is calculated by multiplying the mass and the
gravitational acceleration in the location of the object. Therefore,
the weight of the brass ball on the surface of the Earth is equal to
WE =

mgE = 21.0 kg 9.80 km/s2 = 206 N

while, on the surface of the planet, it has the value


Wp =

mgp = 21.0 kg 12.0 km/s2 = 252 N

End of solution

Page 122 of 261

Chapter 5: problem 34 (Page 132)


Calculate the effective value of g, the acceleration of gravity, above the
Earth's surface at
a. 3200 m.
b. 3200 km.
Solution
The gravitational acceleration (aG) of an object is equal to the force of
gravity acting on the object divided by its mass. Therefore, for an object of
mass m located at a distance r from the center of the Earth, we have

where

mE is the mass of the Earth.

a. The gravitational acceleration of an object placed at an elevation


3200 m above the Earths surface, is calculated as follows:

which is very close to g.


b. If the object is placed at an elevation 3200 km above the Earths
surface, the gravitational acceleration at that height will be

End of solution

Page 123 of 261

Chapter 5: problem 39 (Page 132)

Top

Four 9.5-kg spheres are located at the corners of a square of side 0.60 m.
Calculate the magnitude and direction of the total gravitational force
exerted on one sphere by the other three.
Solution
We may arrange the spheres
such that two of them lie on the
x-axis, as shown in the diagram.
If we then select Sphere 4, we
see that it is pulled towards
Sphere 1 by an attractive
gravitational force with a
magnitude of

where m is the mass of each


sphere and d is the length of
each square side. Similarly,
Sphere 4 is also pulled towards
Sphere 3 by an attractive gravitational force with a magnitude of

To calculate the force of gravity between Sphere 2 and Sphere 4, we need


to find the length of the diagonal of the square. By the Pythagorean
theorem, the diagonal is equal to
. So, Sphere 2 pulls
Sphere 4 by an attractive gravitational force with a magnitude equal to

The resultant force acting on Sphere 4, is therefore equal to the vector sum
of the three forces above. From the symmetry of the problem, we can see
that vertical components of F1 and F3 cancel.
The resultant force is, therefore, directed towards the centre of the square,
and has a magnitude equal to

Page 124 of 261

From the symmetry of the problem you should easily convince yourself that
each sphere experiences a gravitational force that has a magnitude equal
to 3.2 108 N directed towards the centre of the square. You can also do
that by rotating the axes and repeating the arguments above for each
sphere.
End of solution

Chapter 5: problem 44 (Page 132)


The space shuttle releases a satellite into a circular orbit 650 km above the
Earth. How fast must the shuttle be moving (relative to Earth) when the
release occurs?
Solution
For the satellite to perform a uniform circular motion around the Earth, it
must have a velocity such that the required centripetal force is equal to the
force of gravity on the satellite. Mathematically, this is written as

where r is the distance from the satellite to the center of the Earth. In this
equation, the mass of the satellite (mS) cancels out, and the velocity of the
satellite is calculated as follows:

The space shuttle must be rotating around the Earth with this speed
(approximately 27,000 km/h) when the satellite is released.

Page 125 of 261

End of solution

Chapter 5: problem 48 (Page 132)


During an Apollo lunar landing mission, the command module continued to
orbit the Moon at an altitude of about 100 km. How long did it take to go
around the Moon once?
Solution
In this problem, the command module rotates around the moon as a
satellite that moves in a circular orbit of radius
r = rM + 100 km = 1.84 106 m
From the satellite equation

we can derive the following equation for the velocity:

where

mM and rM are the mass and the radius of the moon, respectively.

The time (T) taken by the command module to go around the Moon once is
equal to the distance traveled in one complete rotation divided by the
velocity v. It has the form

To calculate T, we substitute into the equation above as follows:

Notice that the period of rotation is independent of the mass of the


satellite.
End of solution

Page 126 of 261

Chapter 5: problem 54 (Page 132)


A 17.0-kg monkey hangs from a cord suspended from the ceiling of an
elevator. The cord can withstand a tension of 220 N; it breaks as the
elevator accelerates. What was the elevators minimum acceleration
(magnitude and direction)?
Solution
From the free-body diagram, we can see that
two forces are acting on the monkey: the
upward tension of the cord and the downward
force of gravity. The monkeys weight is 167 N,
which is smaller than the maximum tension
sustained by the cord. So, if the elevator does
not accelerate, T becomes equal to the weight
(i.e., T = 167 N), and the monkey will be able to
hang from the cord without breaking it. If the
acceleration is downward, then the net force
acting on the monkey is in that direction too,
which means that T < 167 N. Obviously, the
cord will not break in this case. When the
acceleration is in the upward direction, the
tension in the cord must be greater than the
monkeys weight. According to Newtons second
law, we then write
T

mg = ma

giving the following equation for the upward acceleration

For the cord to break, the tension must become at least 220 N. So, the
minimum acceleration that causes the cord to break is equal to

An upward acceleration greater than 3.14 m/s2 requires tension greater


than 220 N, and the cord would break.
End of solution

Exercise

Page 127 of 261

Solve Problems 30, 31, 33, 41, 43, 53 and 89 at the end of Chapter 5 in
the textbook (pages 131-135).
Answer to 30: 1.62 m/s2

Assignment 2 Questions
Below are the first two questions in Assignment 2. Please keep your
answers until you have answered all the questions of this assignment. Each
question is worth five marks.
1. Three solid spheres of lead, each of mass 9.8 kg, are located at
three corners of a square with side lengths of 50 cm. A small
object is released at the forth corner. Considering only the
gravitational forces among the four objects, determine the
magnitude and direction of the acceleration of the smaller object
when it is released.
2. A spy satellite is in circular orbit around the earth and makes four
revolutions every day.
a. How high above the earths surface is the satellite?
b. Calculate the satellites speed?
Before starting the next lesson, test your comprehension of the material covered in
this unit by turning back to the objectives. Make certain that you can meet each
objective. If you have difficulties, do not hesitate to contact your tutor to discuss the
problem.

Footnote
1

Note that all references to the textbook are to Physics: Principles with
Applications, 6th ed., by Douglas C. Giancoli (Prentice Hall: Upper Saddle River, NJ:
2005).
Top

STUDY GUIDE

Page 128 of 261

Lesson 11
Review I

y completing Lesson 10, you have covered the first part of the course
and should start preparing yourself for the midterm examination. This
lesson contains a quick review of the main concepts covered so far, and
suggests a number practice problems that you can do as a self test. If you
have not done so, please consult your Course Manual for information about
the examination. You can also find a sample exam with a formula sheet and
solutions as well as additional practice problems under the Resources
section of the course.

Scalars and Vectors


A scalar is a physical quantity whose size, or magnitude, is described by a
single number. A physical quantity is said to be a vector if both a numerical
value (called the magnitude) and a direction in space are necessary to
describe the quantity. Displacement, velocity, acceleration and force are
examples of vector quantities. Graphically, a vector is represented by a
directed line segment (or arrow) whose length and direction correspond to
the magnitude and direction of the vector. When making calculations, it is
usually more convenient to decompose a vector into its horizontal and
vertical components. This is especially important when combining different
vectors, because we can simply add similar components to find the
corresponding component of the resultant vector.

Distance and Displacement


There is a distinction between
distance and displacement. By distance we mean the total length traveled
by a body in motion and it is a scalar quantity. Displacement, on the other
hand, is a vector quantity with a magnitude and a direction. The magnitude
of the displacement is the distance between the final and the initial
positions. The direction of the displacement is the direction of the line
segment that joins the initial position to the final position. The diagram on
the right describes the motions of two cars that followed different routes to
move from point a to point b. Notice that even though car 2 traveled a

Page 129 of 261

longer distance than car 1, both


cars have equal displacements.
The distance traveled by a body
is always a positive number. In
one-dimensional motion,
displacement may be expressed
as a positive or a negative
number. The number represents
the magnitude of the
displacement, and the sign (+ or
) indicates the direction of
motion. The accepted convention
is to consider a displacement to
the right of the initial position (or
in the upward direction) as positive and a displacement to the left of the
initial position (or in the downward direction) as negative. However, you
may choose a different convention without changing the reality of the
situation.

Speed and Velocity


Average speed is a scalar quantity defined as the total distance traveled
divided by the elapsed time. Average velocity, on the other hand, is a
vector quantity defined as the displacement divided by the time interval.
Therefore, average velocity has the same direction as the displacement. In
one-dimensional motion, according to the accepted convention, a positive
average velocity will indicate a motion to the right (or upward), and a
negative average velocity will indicate a motion to the left (or downward).
The average velocity during a relatively short time interval (t) is an
approximation of the instantaneous velocity in the middle of that interval.
Making the time interval shorter improves the approximation of the
instantaneous velocity.
The symbol indicates a small change in the value of a quantityx
represents a small change in the value of x, t a small change in the value
of t, and so on. Rememberx is not a product of and x.

Acceleration and Kinematics in One Dimension

Top

Acceleration, defined as the change of velocity over time, does not


necessarily have the same sign as velocity or displacement. If the velocity

Page 130 of 261

of a body and its acceleration are in the same direction (i.e., have the same
sign), the body will speed up and its velocity will increase with time. On the
other hand, if the velocity and the acceleration point in opposite directions
(i.e., have opposite signs), the body will slow down and its velocity will
decrease with time. The table below summarizes these observations.
Sign of Sign of
Description of motion
velocity acceleration
+
+
body accelerates (moving to the right)

body accelerates (moving to the left)


+

body decelerates (moving to the right)

+
body decelerates (moving to the left)
The kinematic equations that describe the motion of a body with constant
acceleration are

where

In horizontal one-dimensional motion, we often write d for distance or


displacement in place of x x0. In vertical one-dimensional motion, we
usually use y0 and y to represent initial and final positions, respectively.
The letter h is often used to represent height.

Freely Falling Objects


Free fall is a special case of motion in one dimension. The acceleration of
the object, in this case, is equal to the gravitational acceleration (g =
9.80 m/s2) and always points in the downward direction. By substituting a
= g in the equations above, we arrive at the following kinematic

Page 131 of 261

equations, appropriate for describing the vertical motion of a freely falling


object:

Projectile Motion
A projectile is a freely falling object with an additional horizontal velocity.
Since the only force affecting the motion of the projectile is the downward
force due to gravity, no acceleration exists in the horizontal direction. As a
result, the horizontal component of the velocity is constant and does not
change with time. The vertical component, on the other hand, behaves
exactly like a freely falling object, moving under gravitational acceleration.
Therefore, it is convenient to analyze the motion of a projectile into its xand y-components. The motion in each direction can then be studied using
the kinematic equations derived for one-dimensional motion. Note that the
velocity vector is always tangent to the path of the projectile.

Newtons Laws of Motion


(1) Newtons First Law of Motion says that, in the absence of a net force
acting on an object, then
z
z

if the object is at rest, it will remain at rest,


if the object is moving, it will continue moving with constant
speed in a straight line.

(2) Newtons Second Law of Motion says the net force F acting on an object
is equal to the mass of the object times its acceleration such that
F=

ma

Because of the vector nature of Newtons second


law, a force acting in a particular direction causes acceleration in that
direction only (i.e., Fx = max, Fy = may etc).
(3) Newtons Third Law of Motion says that for every action (force) there is
an equal, but opposite, reaction.

Page 132 of 261

Normal Force and Apparent Weight


The diagram on the right shows three different
situations of a box resting on a small table. In
the first situation, the box is placed on the table
with no external forces applied. In the second
situation, the box is pressed to the table by
applying a downward external force Fd. In the
third situation, the box is pulled slightly (without
being lifted) by the application of a small
upward external force Fu. The normal force
applied by the table on the box varies, as shown
in the diagrams. In all situations the box is at
rest, which means that the net force acting on it
is zero. Therefore, the normal forces are given
by
Situation 1: FN =

mg

Situation 2: FN =

mg + Fd

Situation 3: FN =

mg Fu

In each of the situations described above, the


normal force is equal to the contact force
between the table and the box. If the table were
replaced the by a weigh scale, the scale would
measure the value of the normal force FN, not
the true weight, mg, of the box. Therefore, in
this case, FN is called the apparent weight of the box.

Static and Kinetic Friction


Friction is a resistance force generated
parallel to the surface of contact of two objects moving, or attempting to
move, against each other. Consider a block resting on a rough surface, as
in the diagram to the right. If a small force is applied to slide the block on
the surface, an equal force of static friction is generated to prevent the
motion. However, the maximum force of static friction generated is equal to
Fs = sFN

Page 133 of 261

where s is the coefficient of static friction.


Thus, if the applied force exceeds Fs, the
net force becomes greater than zero, and
the block starts to slide. When the motion
starts, another type of friction comes into
play, called the force of kinetic friction. This
force is equal to
Fk = kFN
where k is the coefficient of kinetic friction.
Since k is usually smaller than s,
maintaining a motion on a rough surface is often easier that starting the
motion in the first place.

Uniform Circular Motion

Top

The motion of an object traveling with constant speed on a circular path is


called uniform circular motion. The period of the motion is the time it takes
the object to complete one revolution. The speed v of the object is related
to the period T by the equation

where r is the radius of the circular path. Even though the magnitude of the
velocity is constant, the direction changes continuously with time,
indicating continuous acceleration perpendicular to the direction of the
velocity. Since the velocity is tangent to the circular path at all times, the
acceleration always points to the centre of the circular path, thus called
centripetal acceleration. The magnitude of the centripetal acceleration is
equal to

The net force required to generate the centripetal acceleration is called the
centripetal force and has a magnitude equal to

The direction of the centripetal force


is the same as the direction of the centripetal acceleration; that is, towards

Page 134 of 261

the centre of the circular path.


In an unbanked curve, the
centripetal force causing a car to
turn is provided by the force of static
friction between the tires and the
road. In a banked curve, the
centripetal force can be provided
totally by the normal force without
the need for friction. In the figure
above, a car is negotiating a banked
curve on a smooth, frictionless
surface. For the car to stay on track, it should not accelerate either in the
vertical direction or sideways. In other words, the following equations are
satisfied:

Taking the ratio of these two equations, we get the following relation
between the cars velocity, the inclination angle and the radius of a smooth
road:

Law of Universal Gravitation


Two objects of masses

m1 and m2 and separated by a distance r, apply an

attractive gravitational force on each other equal to

where
G = 6.67 1011 N m2/kg2
is the universal constant of gravitation. For example, the force of gravity
applied on the Moon by the Earth is equal to

Page 135 of 261

directed towards the Earth. The Moon applies an equal force on the Earth,
but it is directed towards the Moon. The separation r is equal to the
distance between the centre of the Earth and the centre of the Moon.

Self Test
Below are 18 problems selected from the textbook and grouped into three
sets. To test yourself, consider each set as a practice exam. Give yourself
limited time (approximately two hours) and do not refer to any material
except the formula sheet while working on each set.
Note: The formula sheet is included with the sample midterm exam in the Resources
section.

The purpose of the exam is to test your understanding of the concepts


introduced in the course in Lessons 1-10. The actual exam questions will be
different than the problems below.
Problem Set 1 Chapter
Chapter
Chapter
Chapter
Problem Set 2 Chapter
Chapter
Chapter
Chapter
Problem Set 3 Chapter
Chapter
Chapter
Chapter

2:
3:
4:
5:
2:
3:
4:
5:
2:
3:
4:
5:

problem 23 and 48
problem 53
problem 88
problems 32 and 90
problem 78
problems 16 and 60
problem 69
problems 8 and 67
problem 68
problems 21 and 74
problems 81 and 83
problem 79

Answer to Ch. 2-48: (a) 27 m/s, (b) 37 m, (c) 1.4 s, (d) 4.1 s
Answer to Ch. 2-68: (a) 4.0 s, (b) 24 m
Answer to Ch. 2-78: (a) 5.63 s, (b) 155 m, (c) 55.2 m/s and 60.6 m/s
Answer to Ch. 3-16: 43.3 units
Answer to Ch. 3-60: 0.88 s, 0.95 m

Page 136 of 261

Answer to Ch. 3-74: (a) 13 m/s and 12 m/s, (b) 33 m


Answer to Ch. 4-88: (a,b,c) 75.0 kg, (d) 98.0 kg, (c) 52.0 kg
Answer to Ch. 5-8: 15 m/s
Answer to Ch. 5-32: 16.3 m/s2
Answer to Ch. 5-90: (a) 4.1 104 N, (b) 9.2 105 m
Good luck on your midterm exam!
Top

STUDY GUIDE

Lesson 12
The Work-Energy Principle

he two main concepts in this lesson are work and energy. The terms
are not new to youthey are part of our everyday vocabulary. However,
the terms have definitions and meanings in physics that are different from
those of day-to-day life. In our everyday usage, energy refers to resources
that are capable of producing energyphysical entities, such as natural
gas, oil, coal, etc. Physicists define energy, not as a substance, but as a
property of a physical body or a physical system that is, in essence, an
abstraction. That does not stop us from knowing what energy is. It is
common knowledge that energy comes in different forms. Day in and day
out we get radiant energy from the sun, a benefit that does not cost us a
penny. We also spend huge sums of money to build reactors that provide
nuclear energy. Batteries and household electrical outlets supply electrical
energy, rubber bands store elastic energy, foodstuffs contain caloric
energy, bonfires generate heat energy, and the list goes on. To top it all,
we have Einsteins equation E = mc2 that tells us that mass is also energy.
These examples, however, do not answer the question, What is energy?
Energy can and does change from one type to another. As you drive your
car, heat energy is transformed into mechanical energy. When you honk at
another motorist for disregarding traffic laws, you are converting electrical
energy into sound energy. The electricity that you use in your car comes

Page 137 of 261

from a battery in which chemical energy is converted into electrical energy.


Leaving aside these human-made things, if you look around nature, you
will find examples of the conversion of energy from one form into another.
For example, trees convert the solar energy into chemical energy stored in
the form of wood. However, the knowledge that energy is convertible from
one form into another still does not answer the question What is energy?
As we are unable to define energy in terms of what it is, we define it in
terms of what it does. In most textbooks, energy is defined as the capacity
for doing work. The availability of energy allows work to be done. As
energy changes from one form to another, part of the available energy is
transformed into work. The relationship between work and energy is
embodied in the Work-Energy Theorem, a cornerstone of physics. This
theorem states that work done on a system by an external force increases
the energy of the system, and that energy can be used to do work.
As you might expect, a physicist does not define work in the same way we
do in the context of our everyday lives. If a body moves when a force is
applied to it, a physicist would say that work is done on the body by the
force. According to this definition, if the body does not move, no work is
done , no matter how large the force applied. So, the important thing to
remember is that, in physics, an object must move through a distance for
work to be done. You may spend the whole day trying to move a big piece
of rock from the edge of your yard, pushing and shoving it every which
way, and finally decide that you have done enough work for the day. But, if
the rock has not moved from where it was at the beginning, a physicist
would conclude that you had done no work at all!
By definition, work is the scalar product of two vector quantities, (force and
displacement). That is, it is measured as the product of displacement of an
object and the component of the force acting in the direction of the
displacement. To define work in a more mathematical way, assume that a
constant force F is applied to an object in a certain direction. While the
force is applied, the object makes a displacement d in a direction making
an angle with the direction of the force. We say that work (W) is done on
the object by the force and that the work so done is equal to
W = F d cos
From this equation we see that three conditions must be met before we can
say that work has been done on an object. First, a force (F 0) must be
applied to the object. Second, the object must change its position (d 0).
Third, the displacement must not be perpendicular to the direction of the
applied force ( 90).

Page 138 of 261

Objectives
After completing this lesson, you should be able to
1. define and distinguish between work done by a force and total
work done on a body.
2. define kinetic energy.
3. define potential energy.
4. state the Work-Energy Theorem, and use it to solve problems of
motion.

Reading Assignment
Read Section 6-1 in the textbook.1 You may wish to read Section 6-2 for
interest, but the topic it covers is not considered in this course.

Kinetic Energy
What happens when work is done on a body by an external force?
According to Newtons second law, the external force will affect the
acceleration of the body. If the external force is the only force, the body
will experience acceleration in the direction of the force. If other forces are
present, then the external force will change the net force, which in turn will
change the acceleration of the body. Since the force is also responsible for
the work done on the body, we can conclude that work done on a body by
an external force is related to a change in the velocity of the body.
Let us perform a simple thought
experiment to understand the
process. Consider a body of mass
m moving with constant velocity
v0 on a frictionless surface, as
shown in the diagram. If an
external force F is applied to the body in the direction of motion, the body
will, according to Newtons second law, experience an acceleration a = F/m.
As a result, the bodys velocity will continue to increase until the applied
force is removed. If the body moves a distance d while the force is applied,
its final velocity can be calculated using the kinematic equation

Page 139 of 261

which can be rearranged and written in the form


F

d = mv2 mv02

However, this is nothing but the work done on the body by the applied
force F. It can be seen that the work done on the body by the external
force is equal to the change in the quantity
mass (velocity)2
This quantity, a property of a body in motion, is called the kinetic energy
(KE) of a mass m moving with a velocity v. With this definition, we can
write

mv02 = initial kinetic energy = KEi

mv2 = final kinetic energy = KEf

and
W = KEf KEi = change in kinetic energy
This extremely important relationship is known as the work-energy
principle. It says that work done on a body by an external force is equal to
the change in bodys kinetic energy.
The work-energy principle should be used with caution! When there are
multiple forces acting on a body, then the total work done by all the forces
is equal to the change in the bodys kinetic energy. For example, if the
force of friction between the body and the surface is Ffr, then
W = F d cos 0 = F d = work done by the applied force
Wfr = Ffr d cos 180 = Ffr d = work done by the applied force
Notice that the work done by the force of friction is negative. This is
because the displacement of the body is in the direction opposite to the
resisting fictional force. The net (or total) work done on the body is then
equal to
Wnet = W + Wfr = (F Ffr)d = KEf KEi

Reading Assignment

Page 140 of 261

Read Section 6-3 in the textbook.

Potential Energy

Top

Suppose an object of mass m is at rest on a tabletop. To lift the object, we


need to apply sufficient forcein the upward directionto overcome the
downward force of gravity. The minimum applied force is, therefore, equal
to the weight of the object; that is, F = mg. In this case, the net force
acting on the object is equal to zero, which allows the object to be raised
slowlywithout acceleration.
Assume that the object is raised from its initial position on the tabletop to a
final position at a height h above the tabletop. What is the work done by
the applied force? The displacement vector, in this case, points in the
vertical, upward direction and has a magnitude equal to h. The work done
on the object by the applied force is then equal to
W=F

h cos 0 = mgh

Here, the angle between the displacement and the applied force is zero,
because both vectors point in the upward direction.
What about gravity? Is work done on the object by the force of gravity?
Yes. However, we have to be careful about the sign here, because the force
of gravity and the displacement of the object point in opposite directions,
which results in negative work such that
WG = FG

h cos 180 = mgh

By adding the two equations above


Wnet =

mgh + (mgh) = 0

we see that no (zero) net work is done on the


object, and therefore there is no change in its
kinetic energy. What happened, then, to the
work that was done on the object by the applied
force?
As we will see in the next lesson, the work done
cannot just vanish. In searching for an answer to
this question, let us think of what will happen
when the object is released. As a freely falling
object, it will start accelerating downwards
towards the tabletop and will gain kinetic

Page 141 of 261

energy. Does this gain of kinetic energy have any relation to the work that
was done to lift the object? To answer this question, let us first compute
the final velocity (v) just before the object reaches the tabletop:

The kinetic energy of the object when it reaches the table is equal to
KE =

mv2 = m (2gh) = mgh

which is equal to the work done on the object when it is raised to a height
h above the tabletop.
From the argument above, we see that the work done to raise an object is
not lost. It is returned in the form of kinetic energy when the object is
released back to its original position. So, to put everything together,
scientists came up with an imaginative fiction: that energy can be stored as
potential energyenergy to be used in the future, either as work or for
conversion to other forms of energy. In this scenario, the work done in
lifting the object of mass m through a vertical height h against the force of
gravity is converted into gravitational potential energy,
PE =

mgh

and stays (or is stored) in the general region around the object. This
potential energy can be kept in storage for an indefinite time without any
loss. When we finally let go of the body, the potential energy is converted
into kinetic energy of motion. The work done by the gravitational force
represents a withdrawal from the account of energy that was built up by
the work done in lifting the body.
In dealing with potential energy, we must be aware of some of its
characteristics. Potential energy is always related to a forcegravitational
potential energy arises out of gravitational force, electrical potential energy
comes about only when electrical forces are present, and elastic forces give
rise to elastic potential energy. Thus, potential energy can be defined as
the energy of a body by virtue of its position relative to a force. So, the
potential energy we talk about in this course is different from the energy
stored in a car battery, for example.
Another characteristic of potential energy is that it does not have an
absolute value, like kinetic energy. It is always measured relative to a
reference point, such as the tabletop in the example above. If we imagine
that each object has an account, then the amount of potential energy in the
account when the object is at a particular position is not very important. It

Page 142 of 261

is the amount of energy deposited or withdrawn from the account during


the displacement of an object that is physically meaningful and that can be
related to the work done on the object. In effect, we must realize that
potential energy is not an absolute quantity. When the object is raised from
position A (at the tabletop) to position B, its potential energy increases by
an amount equal to mgh from whatever potential energy it had at A. This
can be made clear by stating that the PE of the body at B, relative to A, is
mgh, or by writing
PE = PEB PEA =

mgh

where PE is understood as the change in the potential energy in moving


the object from A to B.
Gravitational, elastic and electrical forces are all examples of conservative
forces; that is, the work done on an object moving under the action of
these forces is not lost, but is always stored (or conserved) in the form of
potential energy. Frictional forces, on the other hand, are not conservative
forces, and the concept of potential energy cannot be applied to them.

Reading Assignment
Read Sections 6-4 and 6-5 in the textbook.

Chapter 6: problem 3 (Page 162)


A 1300-N crate rests on the floor. How much work is required to move it at
constant speed (a) 4.0 m along the floor against a friction force of 230 N,
and (b) 4.0 m vertically?
Solution
The crate, in this problem, moves at constant speed, with no acceleration.
Therefore, the net force acting on it must be equal to zero. In part (a) the
motion is horizontal and the forces acting on the crate, in this direction, are
the applied force F and the force of friction (Ffr = 320 N). So, net horizontal
force is given by
F Ffr = 0
and the applied force is equal to
F = 230 N
Note that the direction of the displacement is the same as the direction of
the applied force F. Therefore, the work done in pushing the crate a

Page 143 of 261

horizontal distance (d = 4.0 m) along the floor is calculated as follows


W = Fd cos = 230 N 4.0 m cos 0 = 920J
In part (b), the motion of the crate is in the upward direction, and the force
of friction becomes irrelevant. To lift the crate at a constant speed, we will
have to apply an upward force equal to the weight; that is,
F = 1300 N
Since the displacement of the body is d = 4.0 m and it is in the direction of
the applied force, the work done is equal to
W = Fd cos = 1300 N 4.0 m cos 0 = 5200 J
End of solution

Chapter 6: problem 8 (Page 162)

Top

A 330-kg piano slides 3.6 m down a 28 incline and is kept from


accelerating by a man who is pushing back on it parallel to the incline (Fig.
6-36 in the textbook). The effective coefficient of kinetic friction is 0.40.
Calculate:
a. the force exerted by the man.
b. the work done by the man on the piano.
c. the work done by the friction force.
d. the work done by the force of gravity.
e. the net work done on the piano.
Solution
Since the piano does not accelerate, the sum of forces acting on it in any
direction is equal to zero. Therefore, in the direction perpendicular to the
incline, we have
FN

mg cos = 0

resulting in a normal force equal to


FN =

mg cos

The force of kinetic friction acts up the incline to oppose the downward
sliding motion of the piano, and has a magnitude of
Ffr =

kFN = kmg cos

Page 144 of 261

a. From the free-body diagram we


see three forces acting on the
piano in the direction parallel to
the incline. For the sliding
motion to continue down the
incline without acceleration, the
sum of these forces must equal
zero; that is
F+

kmg cos mg sin =

0
The force exerted by the man
is, therefore, equal to

b. Because the piano is sliding down the incline, the displacement is


in the direction opposite that of the applied force (F). Therefore,
the work done by the man on the piano is equal to
Wman = Fd cos 180 = 376.1 N 3.6 m (1) = 1.4 103J
Note that we have used the unrounded value of the applied force.
c. The force of friction is also in the direction opposite the
displacement, and performs a work on the piano equal to

d. The mg cos component of the force of gravity does not do any


work on the piano, since this component of the weight is
perpendicular to the direction of the displacement. The other
component (mg sin ) is in the same direction as the
displacement. Therefore the work done by gravity is equal to

Page 145 of 261

e. Because the forces perpendicular to the incline do not do any


work, the total work done the piano is the summation of the work
done by the forces parallel to the incline. So, we have

End of solution

Chapter 6: problem 18 (Page 162)


How much work must be done to stop a 1250-kg car traveling at
105 km/h?
Solution
The initial speed of the car is equal to

which corresponds to an initial kinetic energy of

When the car comes to a complete stop, its final kinetic energy becomes
zero. So the change in the cars kinetic energy is equal to

According to the work-energy principle, this quantity (5.3 10 J) is


equal to the work done to stop the car. Notice that the work is negative in
this case, because the force applied while the car is slowing down is in the

Page 146 of 261

opposite direction of the displacement.


End of solution

Chapter 6: problem 22 (Page 163)


At an accident scene on a level road, investigators measure a cars skid
mark to be 88 m long. The accident occurred on a rainy day, and the
coefficient of kinetic friction was estimated to be 0.42. Use these data to
determine the speed of the car when the driver slammed on (and locked)
the brakes. (Why does the cars mass not matter?)
Solution
The only horizontal force acting on the car while slowing down is the force
of the kinetic friction between the tires and the road. Since the normal
force, in this case, is equal to the weight of the car, we write
Ffr =

kFN = kmg

The work done on the car by the force of friction during the skid is equal to
W = FNd cos 180 = kmgd
where d is the stopping distance of the car. As we saw in problem 18
above, the work done by friction is equal to the change in the cars kinetic
energy, or W = KE. By substituting, we get
kmgd = 0

mvi2

or

kmgd = mvi2
The mass of the car cancels in this equation and the initial velocity is
calculated to be

It is clear from the equation above that the value of the initial velocity is
independent of the mass. Explanation: If the mass of the car is doubled, its
kinetic energy is also doubled, and therefore, twice the amount of work is
needed to stop it. However, when the mass of the car is doubled, the force
of friction is also doubled, providing the needed work without changing the

Page 147 of 261

stopping distance.
An alternative method for solving this problem involves first calculating the
deceleration of the car

and then using the kinematic equations to calculate the initial velocity. This
is left for you as an exercise.
End of solution

Chapter 6: problem 25 (Page 163)

Top

A 285-kg load is lifted 22.0 m vertically with an acceleration a = 0.160 g by


a single cable. Determine (a) the tension in the cable, (b) the net work
done on the load, (c) the work done by the cable on the load, (d) the work
done by gravity on the load, and (e) the final speed of the load, assuming it
started from rest.
Solution
a. There are two forces acting on the load: its
weight and the upward tension in the cable.
The net force (T mg) is, therefore,
responsible for causing the upward
acceleration of the load. By Newtons
second law, we have
T

mg = ma

So, the tension in the cable is

b. The net work done on the load is equal to net force times the
displacement. This value is calculated as follows

Page 148 of 261

c. The work done by the cable on the load is equal to

d. In computing the work done by gravity on the load, we must note


that the gravitational force and the displacement are in opposite
directions. Therefore, the work done is

e. We can use the Work-Energy Theorem to compute the velocity of


the load after it has moved 22.0 m upwards. Equating the net
work to the final kinetic energy of the load, we get

So, the final speed of the load, assuming it started from rest, is
equal to

End of solution

Chapter 6: problem 30 (page 163)


A 1.60-m tall person lifts a 2.10-kg book from the ground so it is 2.20 m
above the ground.

Page 149 of 261

a. What is the potential energy of the book relative to the ground?


b. What is the potential energy of the book relative to the top of the
persons head?
c. How is the work done by the person related to the answers in
parts a and b?
Solution
Taking the ground level as our reference, we
first calculate the potential energy of the
book at three different heights: h0 = 0,

h1 = 1.60 m, and h2 = 2.20 m.


At the ground level, the potential energy of
the book is, obviously, equal to zero,
because
PE0 =

mgh0 = 2.10 kg 9.80 m/s2 0 = 0

When the book is raised to the level of the persons head, the potential
energy becomes equal to

At a height 2.20 m above the ground, the potential energy increases to

a. When the book is at a height 2.20 m, its potential energy,


relative to the ground, is equal to
PE2 PE0 = 45.3 J
b. Relative to the top of the persons head, the potential energy is
equal to
PE2 PE1 = 12.4 J
c. In this problem, all work done by the person to lift the book is
stored in the form of potential energy. Therefore, the total mount

Page 150 of 261

of work done on the book is equal to 45.3 J. A portion of this


work (32.9 J) was done to raise the book from the ground to the
level of the persons head. The remaining portion (12.4 J) was
done to raise the book from the level of the persons head to a
height 2.20 m above the ground
End of solution

Chapter 6: problem 32 (Page 163)


A spring with k = 53 N/m hangs vertically next to a ruler. The end of the
spring is next to the 15-cm mark on the ruler. If a 2.5-kg mass is now
attached to the end of the spring, where will the end of the spring line up
with the ruler marks?
Solution
Two forces act on the mass when it is attached to the spring: the force of
gravity pulling the mass downward and the elastic force in the spring acting
in the upward direction. The force of gravity is equal to mg, and it does not
change as spring becomes longer. The upward elastic force of the spring,
on the other hand, increases as the spring stretches.
With the right amount of stretch (x) the elastic spring force becomes equal
to the force of gravity, or

kx = mg
In this case, the net force on the mass is
equal to zero. The stretch in the spring
when this happens is calculated as

So, when a 2.5-kg mass is attached to the


spring it stretches down until its end lines
up with the
(15 cm + 46 cm = 61 cm)
mark on the ruler.
End of solution

Page 151 of 261

Exercise
Solve Problems 5, 10, 17, 24, 27 and 29 at the end of Chapter 6 in the
textbook (pages 162-163).
Answer to 10: (a) 1.2 106 J, (b) 3.0 106 J
Answer to 24: 4.41 m

Assignment 2 Questions
3. A hand gun fires a 12.0 g bullet at a speed of 400 m/s.
a. What is its kinetic energy?
b. At what speed must a motorcycle of mass 173 kg move
to have the same kinetic energy as the bullet? Express
your answer in km/h.
4. A dinner plate of mass 510 g is pushed 60 cm along a dining
table by a constant force of 3.0 N directed 22 below the
horizontal. If the coefficient of kinetic friction between the plate
and the tables surface is 0.44, determine the work done on the
plate by
a. the applied force
b. the force of gravity
c. the normal force
d. the force of kinetic friction
Before starting the next lesson, test your comprehension of the material covered in
this unit by turning back to the objectives. Make certain that you can meet each
objective. If you have difficulties, do not hesitate to contact your tutor to discuss the
problem.

Footnote
1

Note that all references to the textbook are to Physics: Principles with
Applications, 6th ed., by Douglas C. Giancoli (Prentice Hall: Upper Saddle River, NJ:
2005).
Top

Page 152 of 261

STUDY GUIDE

Lesson 13
Conservation of Mechanical Energy

n physics, conservation refers to unchangeability, i.e., constancy.


Physicists believe that if a quantity that is characteristic of a system
remains unchanged, while the system itself undergoes change, then the
quantity has some fundamental significance in nature. Physicists have been
successful in finding a number of such conserved quantitiesmass and
energy, momentum, angular momentum, electric charge, to name a few
and they are still finding new ones.
The most celebrated law of physics is The Law of Conservation of
Energy. The popular version states that energy can neither be created nor
destroyed. A physicist says the same thing in a somewhat different way:
the total mechanical energy of an isolated system is a constant, i.e.,
KE + PE = E = Total mechanical energy = constant.
To see how the law works, let us consider projectile motion. Suppose an
object of mass m is projected straight up with a velocity v0 from the
ground. If we take the ground as our reference point for measuring
potential energy, then the total mechanical energy of the object at the time
it is shot up is

As the mass goes up it will lose velocity, thus causing its kinetic energy to
decrease. However, the gravitational potential energy will increase due to
the vertical displacement of the mass against the force of gravity. Suppose
that at a height y above ground, the velocity is v. Then from the equations
of kinematics we can write

v2 = v02 2gy.
Thus the total mechanical energy of the mass at this height is given by

Page 153 of 261

While the object is moving up, its kinetic energy is continuously converted
and stored as potential energy. When all kinetic energy has been converted
into potential energy, the object will be at its maximum height (ymax)
having zero velocity. The total mechanical energy at this instant is equal to

So, we see that the total mechanical energy of the object does not change
as the projectile moves up against gravity. It is always equal to the amount
of energy the projectile had at the start of the motion. When the projectile
starts moving down, its total mechanical energy will also be conserved.
However, in this case, potential energy is converted into kinetic energy,
causing the object to speed up until it hits the ground with the same initial
speed.

Objectives
After completing this lesson, you should be able to
1. state the Law of Conservation of Mechanical Energy.
2. solve problems of motion using the Law of Conservation of
Mechanical Energy.

Reading Assignment
Read sections 6-6, 6-7, 6-8 and 6-9 in the textbook.1

Page 154 of 261

Chapter 6: problem 34 (Page 163)


A novice skier, starting from rest, slides down a frictionless 35.0 incline
whose vertical height is 185 m. How fast is she going when she reaches the
bottom?

Solution
By assuming no energy is lost due to friction or air resistance, the total
mechanical energy of the skier is conserved. In other words, the
summation of potential and kinetic energies remains constant while the
skier is moving down the slope. So, we write
Etop = Ebottom,
or
KEtop + PEtop = KEbottom + PEbottom.
Taking the bottom of the hill as our reference, the initial potential energy of
the skier at the top is equal to PEtop = mgh, where h = 185 m is the height
of the hill, and PEbottom = 0. Since the skier starts from rest at the top of
the hill, then we have KEtop = 0. By substituting into the equation above,
we have
0+

mgh = mv2 + 0,

and the speed of the skier at the bottom of hill is calculated to be

Notice that v is independent of both the skiers mass and the inclination
angle. Assuming ideal skiing conditions, this means that if you begin skiing
with your eight-year-old daughter, you will stay moving together with same
speed and reach the bottom of the slope at the same time.

Page 155 of 261

As an exercise, attempt this problem by calculating first the acceleration of


the skier and then using the kinematics equations to find the final velocity
at the bottom of the incline. You should then appreciate the convenience of
using the Law of conservation of Energy.
End of solution

Chapter 6: problem 38 (Page 163)

Top

A projectile is fired at an upward angle of 45.0 from the top of a 265-m


cliff with a speed of 185 m/s. What will be its speed when it strikes the
ground below? (Use conservation of energy.)
Solution
This is a projectile problem that can be solved using the methods discussed
in Lesson 6. However, we will attempt the problem using the Law of
Conservation of Energy.
Since the projectile moves under the influence of gravity only, its total
mechanical energy is conserved. This means that the mechanical energy of
the projectile when it is fired from the top of the cliff is equal to the
mechanical energy of the projectile just before it strikes the ground, or
KEcliff + PEcliff = KEground + PEground.
By substitution into this equation, we get

m (185 m/s)2 + mg 265 m = mv2 + 0.

In this equation, the mass cancels out and the speed of the projectile when
it strikes the ground is calculated to be

Notice that the final speed of the projectile is independent of the firing
angle! This means that the projectile will always strike the ground below
with the same speed no matter in which direction it is aimed. The
horizontal distance (or range) covered by the projectile depends on the
firing angle.
End of solution

Chapter 6: problem 39 (Page 63)


A vertical spring (ignore its mass), whose spring stiffness constant is

Page 156 of 261

950 N/m, is attached to a table and is compressed down 0.150 m.


a. What upward speed can it give to a 0.30-kg ball when released?
b. How high above its original position (spring compressed) will the
ball fly?

Solution
In this problem there are three types of energy contributing to the total
mechanical energy of the system. In addition to the kinetic and the
gravitational potential energies of the ball, there is also the elastic potential
energy stored in the spring. Therefore, the total mechanical energy of the
system, at any time, is equal to
E = KE + PEgravitational + PEelastic.
To calculate the gravitational potential energy, we need to select a
reference level. For convenience, this is selected to be the level of the
compressed spring as shown in the diagram above. The diagram illustrates
three situations of the system. Situation 1 shows the spring compressed by
x = 0.150 m from its normal length. The kinetic energy is equal to zero in
this situation because the ball is at rest. The gravitational potential energy
is also equal to zero because the ball is located at the reference level. So
the total mechanical energy in this situation is equal to
E1 = 0 + 0 +

kx2.

Page 157 of 261

When the spring is released, it pushes the ball upward, thus accelerating it,
until the spring reaches its normal length (situation 2). In this situation all
the elastic potential energy of the spring has been transformed to the ball
in the form kinetic and gravitational potential energies. So, the total
mechanical energy in this situation is equal to,
E2 =
where

mv2 + mgx + 0,

v is the speed of the ball when it losses contact with the spring.

The ball continues to move upward under the influence of gravity, only
transforming kinetic energy into gravitational potential energy. When the
kinetic energy is depleted the ball will be at its maximum height
(situation 3) just before it starts falling down again towards the spring. In
this situation the total mechanical energy of the system is equal to
E3 = 0 +

mgh + 0.

From the Law of Conservation of Energy the total mechanical energy of the
system is the same in the three situations, i.e.
E1 = E2 = E3.
a. By equating E1 and E2, we have

kx2 = mv2 + mgx,

which allows us to calculate the upward speed of the ball when


released,

b. To find the maximum height of the ball, we equate E1 and E3 as


follows

kx2 = mgh.

The value of h, is then calculated to be

Page 158 of 261

Notice that the second part of this problem can be solved also
using the equations of kinematics for a freely falling object. This
is left for you as an exercise.
End of solution

Chapter 6: problem 40 (Page 163)

Top

A block of mass m slides without friction along the looped track shown in
Fig. 6-39 (in the textbook). If the block is to remain on the track, even at
the top of the circle (whose radius is r), from what minimum height h must
it be released?

Solution
Experience tells us that if the speed of the block is large enough it will
make it around the loop while staying on track. However, for smaller
speeds the block may not make it up the loop and may fall off.
To understand the dynamics of the block in motion, let us analyze the
forces acting on it. In the absence of friction, there are only two forces
acting on the block: the force of gravity (mg) acting in the downward
direction, and the normal force (FN) acting in the direction perpendicular to
the track. So, while moving inside the loop, the normal force on the block
always points towards the center of the circular motion. The centripetal
force acting on the block is, therefore, equal to the normal force plus the
component of the weight perpendicular to the looped track. For the block
remain on track (i.e., move in a circle of radius r), the centripetal force

Page 159 of 261

must always be equal to mv2/r. Notice that


block slows down as it moves up the loop.

v is not constant because the

If the block remains on track and makes it to the top of the loop then it will
also make around the whole loop. At point C both the force of gravity and
the normal force point directly downward towards the center, and the
centripetal force acting on the block is given by

where vc is the velocity of the block at point C. If the block reaches the top
of the loop with relatively high velocity (i.e., mvc2/r>mg) the walls of the
track provide the required normal force so that the equation above is
satisfied. For smaller velocities less normal force is needed, and for the
right velocity (i.e., when mvc2/r = mg) no normal force is needed and the
block negotiates the top of the loop under the force of gravity only. This to
happens when vc is equal to

If the velocity of the block at point C is smaller than this value, then we
have mvc2/rmg. This means that centripetal force (which is equal to mg in
this case) is larger that what is required to cause a circular motion of radius
r. The block, as a result, moves in a smaller circle and loses contact with
the track. Therefore for the block to remain on the track all the way from
to
. The question now is, At what minimum height (h) the block
must be released to reach the top of the loop with this velocity?
Remember that the track is frictionless. This means that the mechanical
energy of the block is conserved and the sum of its kinetic and gravitationa
potential energies does not change while sliding along the track. Comparing
the mechanical energy of the block at points A and C, for example, we have
KEA + PEA = KEC + PEC.
or
0+

mgh = mvc2 + mg(2r).

When the block is just making it around the top of the loop (i.e.,
the relation above becomes

Page 160 of 261

),

mgh = m(rg) + mg(2r).


Both m and g cancel in this equation and the minimum height at which the
block must be released is calculated to be

h = 2.5r.
End of solution

Chapter 6: problem 46 (Page 64)


A cyclist intends to cycle up a 7.8 hill whose vertical height is 150 m.
Assume the mass of bicycle plus cyclist is 75 kg.
a. Calculate how much work must be done against gravity.
b. If each complete revolution of the pedals moves the bike 5.1 m
along its path, calculate the average force that must be exerted
on the pedals tangent to their circular path. Neglect work done by
friction and other losses. Note: The pedals turn in a circle of
diameter 36 cm.
Solution
a. In this problem, we are assuming that the cyclist will cycle up the
hill at a constant speed. In other words, there will be no increase
in the kinetic energy of the bicycle plus cyclist. Also, it is assumed
that no energy will be lost in overcoming the force of friction.
From the work-energy principle, we then conclude that all the
work (W) that needs to be done by the cyclist against gravity is
equal to the increase in the gravitational potential energy of the
bicycle plus cyclist. So, we write

Notice that this is the work to be done by the cyclist, against


gravity. The work done by the force of gravity is the negative of
the value calculated above.
b. To move the bicycle, the cyclist exerts a force (F) on the pedals
tangent to their circular path. When the pedals make one
complete revolution, the applied force acts on the pedals for a

Page 161 of 261

distance equal to the circumference of a circle of radius r =


0.18 m. During this time, the bicycle is elevated by a vertical
distance h1 = 5.1 m sin7.8. Since the work done by the cyclist
is equal to the increase in the gravitational potential energy, we
write
Wone revolution = PEone revolution,
which corresponds to
F 2r =

mgh1.

The average force exerted on the pedals is calculated to be

End of solution
Top

Chapter 6: problem 48 (Page 64)

A 21.7-kg child descends a slide 3.5 m high and reaches the bottom with a
speed of 2.2 m/s. How much thermal energy due to friction was generated
in this process?
Solution
At the top of the slide, just before the child starts moving, her total
mechanical energy is equal to

When the child reaches the bottom of the slide, with velocity
her total mechanical energy is equal to

Page 162 of 261

v = 2.3 m/s,

It is obvious from the numbers above that the child loses most of her
mechanical energy while sliding down. This energy is not actually lost, but
converted into thermal energy due to friction. Therefore, from conservation
of energy, we write
Wfriction = Ebottom Etop = 690 J.
So, 690 J of thermal energy is generated in the process. Notice that the
work done by the force of friction is negative. This is because it acts in the
opposite direction of the childs displacement.
End of solution

Chapter 6: problem 52 (Page 64)


A 110-kg crate, starting from rest, is pulled across a floor with a constant
horizontal force of 350 N. For the first 15 m the floor is frictionless, and for
the next 15 m the coefficient of friction is 0.30. What is the final speed of
the crate?

Solution
One way to solve this problem is to first calculate the acceleration of the
crate in each section of the trip and then substitute into the equations of
kinematics. This is left for you as an exercise.
Another method to solve the problem is to use the Work-Energy Theorem.
Since the motion is restricted to the horizontal plane, no change occurs in
the gravitational potential energy, and the total work done by the
horizontal forces is equal to the change in the crate's kinetic energy.
There are two forces acting on the crate in the horizontal direction. The
constant horizontal force of 350 N acts in the direction of the displacement

Page 163 of 261

during the whole trip of 30 m. So, the work done by this force is equal to
WF = 350 N 30 m = 10,500 J.
The force of friction on the other hand acts in the opposite direction of the
displacement during the second half of the trip for 15 m only. So, the work
done by friction is equal to

Since the total work on the crate is equal to the change in its kinetic
energy, we write
WF + Wfriction = KE =
where

mv2,

v is the final speed of the crate. Solving for v, we get

This is the final speed of the crate.


End of solution

Chapter 6: problem 86 (Page 66)


A 65-kg student runs at 5.0 m/s, grabs a rope, and swings out over a lake
(Fig. 6-45 in the textbook). He releases the rope when his velocity is zero.
a. What is the angle when he releases the rope?
b. What is the tension in the rope just before he releases it?
c. What is the maximum tension in the rope?
Solution
a. The swinging motion starts when the student grabs the rope
while running at 5.0 m/s. As the rope swings away from the
vertical position, the student loses speed while being elevated

Page 164 of 261

above the surface of the lake. Since friction and air resistance are
ignored, the total mechanical energy is conserved in this
problem. This means that kinetic energy of the student at the
beginning of the swing is equal to the increase in his gravitational
potential energy by the end of the swing, or

mv2 = mgh.

From this equation we can calculate the maximum increase in


elevation,

which (according to the diagram below) corresponds to an angle

When the rope makes this angle with the vertical, the student
releases the rope.

b. The diagram above shows a free-body diagram of the student


when he is at the farthest point of the swing. Since no motion is

Page 165 of 261

allowed in the direction parallel to the rope, the net force acting
on the student in this direction must be equal to zero, or
T

mg cos = 0.

The tension in the rope just before the student releases is then
equal to

c. The maximum tension in the rope occurs at the beginning of the


swing. In this position it is not enough for the rope to provide
tension equal to the students weight. Extra tension is required to
act as the centripetal force, allowing the student to rotate in an
arc of radius 10.0 m. So, in this position, we have

The maximum tension in the rope is thus equal to

End of solution

Exercise
Solve Problems 35, 37, 43, 51, 53, 55 and 77 at the end of Chapter 6 in
the textbook (pages 163-165).

Assignment 2 Questions
5. A 15 kg block slides along a horizontal frictionless surface 3.0 m
above the ground at a constant speed of 2.0 m/s. The block then
slides down an incline that makes an angle of 35 with the
horizontal and has a coefficient of kinetic friction equal to 0.30.
After reaching the end of the incline, the block continues sliding

Page 166 of 261

horizontally across the frictionless ground. Calculate the kinetic


energy of the block as it slides
a. along the upper surface.
b. along the ground.

6. A block of mass 2.5 kg is placed against a compressed spring


(k = 2900 N/m) at the bottom of an inclined plane, as shown.
When the spring is released the block is projected up the incline
and the spring expands by 14 cm to its normal length.
a. Calculate the maximum distance traveled by the block
up the incline without friction.
b. Repeat the calculation with friction, taking

k = 0.22.

Before starting the next lesson, test your comprehension of the material covered in
this unit by turning back to the objectives. Make certain that you can meet each
objective. If you have difficulties, do not hesitate to contact your tutor to discuss the
problem.

Footnote
1

Note that all references to the textbook are to Physics: Principles with
Applications, 6th ed., by Douglas C. Giancoli (Prentice Hall: Upper Saddle River, NJ:
2005).
Top

STUDY GUIDE

Page 167 of 261

Lesson 14
Linear Momentum and its Conservation

n this lesson, we will introduce one of the most important concepts in


physics. The new quantity associated with this concept is called linear
momentum (p). It is simply defined as the product of mass and velocity.
So, the linear momentum of a particle of mass m moving with velocity v is
given by

p = mv.
To see how the concept of linear momentum (commonly referred to as
momentum) fits into the Newtonian scheme, we will take another look at
Newtons Second Law of Motion. The mathematical form of the law (F =
ma) can be rewritten as

where we have made use of the fact that the mass is a constant quantity.
With the notation p = mv for linear momentum the second law takes the
form

The equation above is true for motion in one dimension. In general,


velocity is a vector quantity. Since linear momentum is defined as the
product of mass and velocity, it must also be a vector, since multiplication
of a vector (v) by a scalar (m) results in a vector (p). With the definition
p=

m v,

Newtons Second Law takes the form

Newtons Second Law takes on a new interpretation in light of the new

Page 168 of 261

relationship. An unbalanced force acting on a body, for a period of time t,


causes a change in the bodys momentum equal to F t. We can also say
that the rate of change of momentum is equal to the net force acting on
the body. In Newtonian mechanics, the two statements, F = ma and F =
p/t, are equivalent and interchangeable.

Objectives
After completing this lesson, you should be able to
1. define linear momentum of a body and linear momentum of a
system of bodies.
2. reinterpret Newtons Second Law as a relationship between force
and the rate of change of linear momentum.
3. state the Law of Conservation of Linear Momentum.

Reading Assignment
Read section 7-1 in the textbook.1

The Law of Conservation of Linear Momentum

Top

From Newtons second law, F = p/t, it is evident that in absence of a


force, i.e., when F = 0, the rate of change of momentum is zero. In other
words the momentum remains constant. The generalization of this
relationship, which states that the total momentum of an isolated system is
constant, is the celebrated Law of Conservation of Linear Momentum.
Consider the simple case of a single object moving in outer space. If there
are no forces acting on the object, we know that its acceleration is zero and
the velocity remains constant. Since the mass is a constant quantity, then
the momentum (mv) is also a constant quantity.
Let us now consider a system (or a collection) of particles interacting in
space. In this case we need to differentiate between two types of forces
acting on the particles. The first type is the internal forces, which are
applied by the different particles in the system on each other. Examples
include gravitational attractions between the particles, elastic forces due to
collisions, and electrical forces if the particles are charged. The second type
of forces acting on the particles is the external forces, which are applied on
the particles by sources that exist outside the system. For example, the

Page 169 of 261

system could be in the vicinity of a large planet and is affected by its


gravitational field.
Particles of different masses may collide with each other within an isolated
system and may change velocities, resulting in changes in the momentum
of the individual particles. However, the total linear momentum of the
particles within the system remains a constant as long as no external forces
act on the system. Suppose a system consists of the masses m1, m2, m3,
. . . etc. with velocities v1, v2, v3, . . . etc. The Law of Conservation of
Linear Momentum states that the total momentum p of the system, i.e.,
the vector sum of the momenta p1 = m1 v1, p2 = m2 v2, p3 = m3 v3, . . .
etc., is a constant. Or, we can write
p = p1 + p2 + p3 + . . . etc. = constant.
Even, if at a later time it is found that the momenta of the particles within
the isolated system have changed to p1 = m1v1, p1 = m1v1, p1 = m1v
1 . . . etc., the total momentum of the system will remain constant such

that
p = p1 + p2 + p3 + . . . etc.
= p1 + p2 + p3 + . . . etc. = p = constant
This equation is the mathematical statement of the Law of Conservation of
Linear Momentum for an isolated system. As is the Law of Conservation of
Energy, the Law of Conservation of Linear Momentum is a cornerstone of
physics; together, the two form a powerful tool for analyzing motion.

Reading Assignment
Read sections 7-2 and 7-3 in the textbook.

Chapter 7: problem 2 (Page 188)


A constant friction force of 25 N acts on a 65-kg skier for 20 s. What is the
skier's change in velocity?
Solution
One way to solve this problem is to first calculate the deceleration of the
skier caused by the force of friction, and then substitute into the kinematic
equation v = v0 + at to find the change in velocity.

Page 170 of 261

The problem can also be solved using the form of Newtons Second Law
introduced in this lesson. This is because the applied force and the duration
of the applied force are known. If the skier is considered to be moving in
the positive direction, the force of friction acting on her becomes negative,
and the change in momentum is calculated to be
p = F t = 25 N 20 s = 500 N s.
Since momentum is defined as mass times velocity, we can write
p =

m v.

The change in the skiers velocity is then calculated to be

where the negative sign indicates that the velocity decreases as expected.
End of solution

Chapter 7: problem 4 (Page 188)


A child in a boat throws a 6.40-kg package out horizontally with a speed of
10.0 m/s (Fig. 7-31 in the textbook). Calculate the velocity of the boat
immediately after, assuming it was initially at rest. The mass of the child is
26.0 kg, and that of the boat, 45.0 kg. Ignore water resistance.
Solution
The system in this problem consists of the boat, the child and the package.
Before the child throws the package, the boat is assumed to be standing
still. Therefore, the initial momentum of the system is equal to zero. For
the child to throw the package, he must apply a force on it. According to
Newtons Third Law, an equal reaction force is also applied on the child and
the boat. However, since these are internal forces, the momentum of the
system is conserved. This means that the sum of the momenta of the boat,
the child and the package remains zero even when the package is moving
away from the boat. So, we write

Solving for v, we get

Page 171 of 261

So, the resulting velocity of the boat carrying the child, after the child
throws the package, is 0.901 m/s in the direction opposite to that in which
the child threw the package.
End of solution

Chapter 7: problem 9 (Page 188)


During a Chicago storm, winds can whip horizontally at speeds of
100 km/h. If the air strikes a person at the rate of 40 kg/s per square
meter and is brought to rest, estimate the force of the wind on a person.
Assume the person is 1.50 m tall and 0.50 m wide. Compare to the typical
maximum force of friction ( 1.0) between the person and the ground, if
the person has a mass of 70 kg.
Solution
The rate at which the air strikes the person each second is 40 kg/m2 and
the area of the person is 1.5 m 0.50 m = 0.75 m2. Therefore, the mass
of the air striking the person each second is

mair = 40 kg/m2 0.75 m2 = 30 kg.


As stated in the problem, the air loses all its speed (v = 100 km/h =
27.78 m/s) when it strikes the persons body. This means that the
momentum of the air is transferred to the person, who gains momentum
equal to
p =

mairv = 30 kg 27.78 m/s = 830 kg m/s

during a period of one second. According to Newtons Second Law, the


average force applied on the person due to the wind is equal to

This is greater than the maximum frictional force between the person and
the ground
Ffr =

mg = 1.0 70 kg 9.80 m/s2 = 690 N.

End of solution

Page 172 of 261

Chapter 7: problem 10 (Page 88)

Top

A 3800-kg open railroad car coasts along with a constant speed of 8.60 m/s
on a level track. Snow begins to fall vertically and fills the car at a rate of
3.50 kg/min. Ignoring friction with the tracks, what is the speed of the car
after 90.0 min?

Solution
The system of interest in this problem consists of the railroad car and the
falling snow. While the snow is falling down, its horizontal velocity is equal
to zero. So, the initial horizontal momentum of the system is equal to that
of the moving car. Since no external forces are involved, the momentum of
the system is conserved. This means that the final momentum of the
system (car filled with snow) is the same as the initial momentum. It is
true that the snow has an initial velocity in the downward direction.
However, this will only affect the vertical momentum of the system which is
of no concern to us in this problem.
The amount of snow that falls on the car during a period of 90.0 min is
equal to

ms = 3.50 kg/min 90.0 min = 315 kg.


The initial horizontal momentum of the system is equal to

pi = mcvc,
where

mc is the mass of the railroad car and vc is its initial speed. The final

horizontal momentum after 90.0 min of continuous snowfall is equal to

pf = (mc + ms)v.
Since pi = pf, then we write

mcvc = (mc + ms)v,


and the final speed of the car after 90.0 min is equal to

Page 173 of 261

This reduction in speed can be explained as follows: When the snow


reaches the car, a force is applied on it and accelerates until its velocity
becomes equal to that of the car. During this time, an equal reaction force
also acts on the car in the opposite direction, causing it to decelerate.
However, due to the large difference in mass, the deceleration of the 3800kg car is much smaller than the acceleration experienced by the 315 kg of
snow.
End of solution

Chapter 7: problem 13 (Page 88)


A 975-kg, two-stage rocket is traveling at a speed of 5.80 103 m/s with
respect to Earth when a pre-designed explosion separates the rocket into
two sections of equal mass that then move at a speed of 2.20 103 m/s
relative to each other along the original line of motion.
a. What are the speed and direction of each section (relative to
Earth) after the explosion?
b. How much energy was supplied by the explosion? [Hint: What is
the change in KE as a result of the explosion?]
Solution

m be the original mass of the rocket, and V its velocity. Let v1


and v2 be the velocities of the two sections of the rocket, each of
mass m/2, after the main rocket separates into two pieces. The

a. Let

system in this case is the whole rocket including all its sections.
Since the forces created by the explosion are internal forces, then
the total momentum of the system is conserved. In other words,
the initial momentum of the rocket before the explosion is equal
to the total momentum of the all sections of the rocket after the
explosion. So, we write

mV = (m/2) v1 + (m/2) v2,


which leads to the following relation between v1 and v2

v1 + v2 = 1.16 104 m/s

Page 174 of 261

The given relative velocity of the two separated sections gives


another relation between v1 and v2

v1 v2 = 2.20 103 m/s


By adding the two equations above, we get

Putting in the value of v1 in the


second equation, we find that the
velocity of the second section is
equal to

v2 = v1 2.20 103 m/s =


4.70 103 m/s.

Notice that the two sections


continue moving forward. The
force created by the explosion
accelerated the first section and
decelerated the second one.
b. To find the energy supplied by the explosion, we need to compute
the difference in the kinetic energies before and after the
explosion. The kinetic energy of the system before the explosion
is

After the explosion, the total kinetic energy of the two smaller
rockets is

Page 175 of 261

The difference in kinetic energies is

This is the amount of energy supplied by the explosion.


End of solution

Chapter 7: problem 16 (Page 188)


A 12-kg hammer strikes a nail at a velocity of 8.5 m/s and comes to rest in
a time interval of 8.0 ms.
a. What is the impulse given to the nail?
b. What is the average force acting on the nail?
Solution
When a force F acts on an object during a time interval t, then an impulse
equal to
Impulse = F t
is given to the object. However, since F t = p, then the impulse is also
equal to the change in the objects momentum.
a. When it hits the nail and comes to rest, the hammer loses all its
initial momentum. However, the total momentum of the system
(hammer and nail) is conserved, due to the absence of external
forces. Therefore, the decrease in the hammers momentum is
equal to the increase in the momentum of the nail, and the
impulse given to the nail is equal to

Page 176 of 261

b. From the equation that gives the definition of the impulse, we see
that the average force acting on the nail is equal to

End of solution

Exercise
Solve Problems 1, 7, 11, 12 and 17 at the end of Chapter 7 in the textbook
(page 188).
Answer to 12: 0.69 m/s

Assignment 2 Questions
7. A 15,000 kg loader traveling east at 20 km/h turns south and
travels at 25 km/h. Calculate the change in the loaders
a. kinetic energy.
b. linear momentum.
8. A 470-g firework is traveling straight up at 13 m/s when it
explodes into two pieces. The smaller piece (150 g) shoots off
horizontally towards the East at 20 m/s. Find the speed and
direction of the other piece directly after the explosion.
Before starting the next lesson, test your comprehension of the material covered in
this unit by turning back to the objectives. Make certain that you can meet each
objective. If you have difficulties, do not hesitate to contact your tutor to discuss the
problem.

Footnote

Page 177 of 261

Note that all references to the textbook are to Physics: Principles with
Applications, 6th ed., by Douglas C. Giancoli (Prentice Hall: Upper Saddle River, NJ:
2005).
Top

STUDY GUIDE

Lesson 15
Collisions

ollision is a very important subject in physics. Physicists perform


collision experiments in the advanced research of nuclear and particle
physics. In such experiments, highly energetic particles (electrons, protons
or atoms nuclei) are smashed together for the purpose of studying the
structure of matter at the subatomic level and understanding the
fundamental interactions between the elementary particles in nature. Such
experiments, which are done at the microscopic level, require highly
sophisticated equipment and complex theoretical calculations due to the
complications of relativistic and quantum effects. In this lesson, we will
study collisions at the macroscopic level using only the principles of
Newtonian mechanics.
Consider two objects that are moving and approaching each other in space.
In the absence of any external forces, the two objects form an isolated
system. When the two objects collide, they apply forces on each other
during the short period of the collision. These are internal forces that will
affect the momentum of each object, but not the total momentum of the
system. Therefore, if one object gains momentum as a result of the
collision, then the other object will lose momentum by exactly the same
amount. In other words, the total momentum of both objects before the
collision is equal to the total momentum after the collision. Thus, linear
momentum is conserved in any collision.
Collisions are divided into two classeselastic and inelasticbased on
kinetic energy considerations. Total energy and total momentum are
always conserved in a collision, but if kinetic energy is also conserved, the
collision is called elastic.

Page 178 of 261

Objectives
After completing this lesson, you should be able to
1. define the relationship between impulse and momentum.
2. solve elastic and inelastic collision problems in one dimension.
3. solve elastic and inelastic collision problems in two dimensions.

Reading Assignment
Read section 7-4 in the textbook.1

Collisions in One Dimension


Consider two masses,

m1 and

m2, moving on the same line and


approaching each other with
velocities v1 and v2. Suppose
that the two masses collided with
each other head-on and bounced
with velocities v1 and v2 such that both stay on the line, as shown in the
diagram. It follows from conservation of momentum that

m1v1 + m2v2 = m1v1 + m2v2


This equation applies to all kinds of collisions, and it is always true, as long
as no external forces are present. The total energy is always conserved in
any collision. If we notice that the kinetic energy is conserved along with
the total energy, then the collision is elastic. In this case, the following
equation is also satisfied:

m1v12 + m2v22 = m1v12 + m2v22.

These two equations, the conservation of momentum and the conservation


of kinetic energy, describe a perfectly elastic collision. If kinetic energy is
not conserved, then only the conservation of momentum equation applies,
and the collision is called inelastic.
Notice that in the conservation of momentum equation, the velocity
appears as a vector quantity. However, for motion in one dimension we

Page 179 of 261

only have two directions (e.g., right or left). Therefore, it is usually


convenient to assume positive velocities for objects moving to the right and
negative velocities for objects moving to the left.

Reading Assignment
Read sections 7-5 and 7-6 in the textbook.

Collisions in Two Dimensions


If, after the collision, the two masses are
scattered in directions that are not along
the original line of motion, then we have
a collision in two dimensions. The law of
conservation of momentum also applies
in this case. However, we need to be
careful to pay close attention to the
vector nature of velocity and momentum.
To be specific, both the horizontal and
the vertical components of momentum
are separately conserved. So, the
equation representing conservation of
momentum is resolved into two scalar
equations along the mutually
perpendicular x- and y-axes as follows

m1v1x + m2v2x = m1v1x + m2v2x,


m1v1y + m2v2y = m1v1y + m2v2y.
The equation representing conservation of kinetic energy, in the case of
elastic collisions, will be the same as that used for collisions in one
dimension, i.e.,

m1v12 + m2v22 = m1v12 + m2v22.

This is because only the magnitude of the velocity enters into the equation.

Reading Assignment
Read section 7-7 in the textbook.

Page 180 of 261

Chapter 7: problem 24 (Page 189)

Top

Two billiard balls of equal mass undergo a perfectly elastic head-on


collision. If one balls initial speed was 2.00 m/s, and the other's was 3.00
m/s in the opposite direction, what will be their speeds after the collision?
Solution
The collision in this problem is
head-on, which means that the
two balls continue to move along
the same line (in one dimension)
after the collision. Since no
external forces are involved, the
total momentum of the system
(balls A and B) is conserved. So,
we write

mvA + mvB = mvA + mvB,


where m is the mass of each ball. The problem mentions that the collision
between the two billiard balls is elastic. This means that the kinetic energy
of the system is also conserved, and we can write

mv2A + mv2B = mvA2 + mvB2.

By canceling common factors, the two equations above reduce to

vA + vB = vA + vB,
vA2 + vB2 = vA2 + vB2.
Here we have two equations with two unknowns (vA, and vB) and this
becomes a matter of solving these two equations simultaneously. As a first
step, we rearrange the terms in the two equations above as follows:

vA vA = vB vB,
vA2 vA2 = vB2 vB2.
The algebraic identity (x2 y2) = (x y)(x + y), allows us to factor the
second equation above, such that we have

vA vA = vB vB,

Page 181 of 261

(vA vA) (vA + vA) = (vB vB) (vB + vB).


The two terms (vA vA) and (vB vB) cancel out in the second equation.
This is because, according the first equation, these two terms are equal.
So, the two equations above reduce to

vA vA = vB vB,
vA + vA = vB + vB.
A simple way to solve these two equations is to add them together and
then subtract them. If we do that, we get the following two relations:
2vA = 2vB,
2vA = 2vB,
which give the following solution

vB = vA,
vA = vB.
Since the initial velocities of the balls are vA = 2.00 m/s and vB = 3.00
m/s, then after the collision, the velocities become

vA = 3.00 m/s,
and

vB = 2.00 m/s.
So, the two balls simply exchanged velocities. Note that this general result
applies only to elastic head-on collisions between two objects of equal
mass.
End of Solution

Chapter 7: problem 29 (Page 189)


In a physics lab, a cube slides down a frictionless incline as shown in Fig. 735 (in the textbook) and elastically strikes another cube, that is only onehalf its mass, at the bottom. If the incline is 30 cm high and the table is 90
cm off the floor, where does each cube land? [Hint: Both leave the incline
moving horizontally.]

Page 182 of 261

Solution
The solution of this problem can be divided into three steps. In the first
step, we find the velocity of cube A, just before it strikes the other cube at
the end of the incline. In the second step, we calculate the velocities of the
cubes just after the collision. The last step of the solution is a projectile
problem in which we calculate the horizontal distance moved by each cube
before landing on the floor.
Step 1: When it slides down the frictionless incline, the gravitational
potential energy of cube A is transformed into kinetic energy, such that

mAgh = mAv2A.
So, the velocity of cube A just before it strikes the other cube at the end of
the incline is equal to

Step 2: The two blocks collide elastically at the end of the incline, which
means that both the linear momentum and the kinetic energy of the
system are conserved. From the information we have about the collision

Page 183 of 261

we write the following equations, representing conservation of linear


momentum and kinetic energy, respectively:

mAvA + mBvB = mAvA + mBvB,


and

mAv2A + mBv2B = mAvA2 + mBvB2.

If we substitute vB = 0 and

mA = 2mB, the two equations above reduce to

2vA = 2vA + vB,


2v2A = 2vA2 + vB2.
By rearranging terms, we get

vB = 2(vA vA),
vB2 = 2(v2A vA2).
One way to solve these two equations is to substitute vB from the first
equation into the second one as follows:
2(vA vA)2 = (v2A vA2).
After factorizing the two terms, the above equation becomes
2(vA vA) (vA vA) = (vA vA) (vA + vA),
or
2(vA vA) = (vA + vA).
Solving for vA, we get

and vB is calculated by substituting back as follows

The final velocities of the two blocks directly after the collision are,

Page 184 of 261

therefore, equal to

vA = 0.808 m/s
and

vB = 3.233 m/s
Step 3: After the collision, the two blocks leave the incline moving in the
horizontal direction with the velocities calculated above. So, initially the
blocks start the projectile motion at the same height above the floor with
zero initial vertical velocity. Both blocks, therefore, reach the floor at the
same time, which is calculated by substituting into the kinematic equation
= y0 + vy0t +

at , as follows:

0 = 0.90m + 0 + (9.80 m/s2) t2,


where

t is found to be

t = 0.428s.
During this time, block A moves a horizontal distance equal to

xA = vAt = 0.808 m/s 0.428 s = 0.35 m,


while block B moves a larger horizontal distance equal to

xB = vBt = 3.233 m/s 0.428 s = 1.4 m.


End of Solution

Chapter 7: problem 35 (Page 190)

Top

A 920-kg sports car collides into the rear end of a 2300-kg SUV stopped at
a red light. The bumpers lock, the brakes lock, and the two cars skid
forward 2.8 m before stopping. The police officer, knowing that the
coefficient of kinetic friction between tires and road is 0.80, calculates the
speed of the sports car at impact. What was that speed?
Solution
Since the collision in this problem is inelastic, no conservation of kinetic
energy can be assumed. Also, the equation representing conservation of
momentum is not sufficient to provide a direct solution to the problem. This
is because the initial and the final speeds are unknowns. So, to find the
initial speed of the sports car using the conservation of momentum

Page 185 of 261

equation, we
need to figure
out the
combined
speed of the
two vehicles
directly after
the collision.
The
information we
have about the
motion of the
two cars after
the collision is very useful in this case. This is because the work done by
the force of friction to decelerate and then stop the two cars is equal to the
kinetic energy directly after the collision. This is written as
(msc +

msuv)v2 = (msc + msuv)g d,

msc and msuv are the masses of the sports car and the SUV
respectively, and d is the stopping the distance of the two cars. The masses

where

in the equation above cancel out and the velocity of the cars just when the
bumpers lock is calculated to be

Knowing the velocity just after the collision, we can now write the
conservation of momentum equation,

mscvsc + msuvvsuv = (msc + msuv)v,


and then substitute as follows
920 kg vsc + 0 = (920 kg + 2300 kg) 6.626 m/s.
The speed of the sports car at impact is, therefore, equal to

End of Solution

Chapter 7: problem 40 (Page 190)

Page 186 of 261

A radioactive nucleus at rest decays into a second nucleus, an electron, and


a neutrino. The electron and neutrino are emitted at right angles and have
momenta of 9.30 1023 kgm/s and 5.40 1023 kgm/s, respectively.
What are the magnitude and direction of the momentum of the second
(recoiling) nucleus?
Solution
The system in this problem consists of a single particle at rest that decays
into three other particles moving with different speeds in different
directions. The decay of the original particle occurs due to internal
interactions only. So, the linear momentum of the system is conserved.
Since the initial momentum is equal to zero, the total momentum of the
system after the decay is also equal to zero. In other words, the vector
sum of the momenta of the electron, the neutrino and the recoiling nucleus
is equal to zero, or
Pn + Pe + Pv = 0.
Since this is a two-dimensional
problem, we should first set our
coordinates and then resolve the
momentum into x and y
components. As a convention
(see the diagram above) we
choose the direction of motion of
the neutrino as the positive
direction of the y-axis and the
direction of motion of the
electron as the negative direction
of the x-axis. According to this
convention, we have

Since each component of the momentum is individually conserved, the


conservation of momentum equation is replaced by the following two scalar
equations corresponding to x and y directions:

Pnx + Pex + Pvx = 0,


and

Page 187 of 261

Pny + Pey + Pvy = 0.


By substitution into these two equations, we get the x and the y
components of the recoiling nucleus momentum, i.e.,

Pnx = 9.30 1023kgm/s


Pny=5.40 1023kgm/s.
The magnitude of the recoiling nucleus momentum is, therefore, equal to

which makes, according to the diagram above, an angle

below the positive direction of the x-axis. To make the answer independent
of our choice of the coordinates, we can say that the recoiling nucleus
makes an angle of approximately 150 with the direction of motion of the
electron.
End of Solution

Chapter 7: problem 41 (Page 190)


An eagle (mA = 4.3 kg) moving with speed vA = 7.8 m/s is on a collision
course with a second eagle (mB = 5.6 kg) moving at vB = 10.2 m/s in a
direction perpendicular to the first. After they collide, they hold onto one
another. In what direction, and with what speed, are they moving after the
collision?
Solution
Because the two eagles hold onto one another and move as a single body,
their collision is inelastic and kinetic energy is not conserved. Therefore,
only the conservation of linear momentum can be used in the solution of
this problem. So, if pA and pB are the initial momenta of the two eagles
before the collision and p is the final momentum of both eagles after the

Page 188 of 261

collision, then
from
conservation of
momentum we
have
PA + PB = p
or

mAvA + mBvB
= (mA + mB)v
Notice that the
symbols p and
v are
capitalized to
represent
vector
quantities.
Since the problem is two-dimensional, the equation above is resolved into
two equations, corresponding to the x and y directions, as follows

mAvAx + mBvBx = (mA + mB)vx,


mAvAy + mBvBy = (mA + mB)vy.
For convenience, we will consider the direction of motion of eagle A as the
positive x-direction and the direction of motion of eagle B as moving along
the positive y-axis, as shown in the diagram above. After the collision, the
two eagles move together in a direction making an angle with the
positive x-axis. So, in this coordinate system, the x and y components of
the velocities involved are

Substituting back into the conservation of momentum equations, we get


4.3 kg 7.8 m/s + 0 = 9.9 kg vcos ,

Page 189 of 261

0 + 5.6 kg 10.2 m/s = 9.9 kg vsin .


If we divide the second equation by the first, v cancels out and we get

The angle is then calculated to be


= tan1(1.703) = 60.
Next, we substitute the calculated back into the x component of the
conservation of momentum equation such that
4.3 kg 7.8 m/s = 9.9 kg vcos 59.58.
Solving for v, we get

So, the two eagles will move with a speed of 6.7 m/s in the direction
making an angle of 60 with the direction of flight of eagle A.
End of Solution

Chapter 7: problem 44 (Page 190)

Top

Two billiard balls of equal mass move at right angles and meet at the origin
of an x y coordinate system. Ball A is moving upward along the y axis at 2.0
m/s, and ball B is moving to the right along the x axis at 3.7m/s. After the
collision, assumed elastic, ball B is moving along the positive y axis (Fig. 737 in the textbook). What is the final direction of ball A and what are the
two speeds of balls A and B?
Solution
Let us assume that ball A
bounces in a direction that makes an angle with the positive x-axis (see
the diagram to the right). The goal is to find the final velocity (vB) of ball
B, and the final velocity (vA) and direction () of ball A after the collision.
In the problem, it is mentioned that the collision between the two billiard
balls is elastic. This means that both linear momentum and kinetic energy
are conserved in the collision. Converting this statement into equations, we
write

Page 190 of 261

mvB = mvA cos,


mvA = mvB + mvAsin,
and

mv2A + mv2B = mvA2

mvB2,

which correspond to
conservation of momentum
along the x-axis, conservation
of momentum along the y-axis
and conservation of kinetic
energy, respectively. Since the
two balls have equal mass, m cancels out, and the three equations above
reduce to

vB = vA cos ,
vA = vB + vA sin ,
and

vA2 + vB2 = vA2 + vB2.


So, we now have three equations that need to be solved, simultaneously,
to find the three unknowns vA,vB and .
Rearranging terms, the three equations above become

vB = vAcos,
vA vB = vAsin,
and

vA2 vB2 = vA2 vB2.


The purpose of this rearrangement is to get rid of the angle and reduce
this set of equations to two equations with two unknowns. To do that, we
square the first two equations (which correspond to conservation of
momentum), and then add them together. As a result, we get

Page 191 of 261

where we have used of the trigonometric identity cos2 + sin2 = 1. With


this merging of the first two equations, the set of equations that we need to
solve reduces to

vB2 + (vA vB)2 = vA2


and

vA2 vB2 = vA2 vB2.


Further rearrangement of terms in the equations above become
(vA vB)2 = vA2 vB2
and

vA2 vB2 = vA2 vB2.


You can now see clearly that if we subtract these two equations, we will get
rid of vA and end up with one equation with one unknown,
(vA vB)2 (vA2 vB2) = 0.
To solve for vB we first expand the first term as follows:
(vA2 2vAvB + vB2) (vA2 vB2) = 0.
then, we subtract the two terms to get

vB(vB vA) = 0.
For this equation to be satisfied, either vB is equal to zero or vB is equal to

vA. However, the problem mentions that ball B continues to move along the
positive y-axis after the collision. So, vB is greater than zero and we must
have

vB = vA.
To find the speed of ball A after the collision, we substitute for vB into the

Page 192 of 261

conservation of kinetic energy equation such that

vA2 + vB2 = vA2 + vA2,


which gives

vA = vB.
If we now substitute into the equation corresponding to conservation of
momentum along the x-axis, we get

So, the values of the three unknown parameters in this problem are

vB = vA = 2.0 m/s,
vB = vA = 2.0 m/s
and
= cos1(1) = 0.
Notice that the two billiard balls simply exchange velocities.
End of Solution

Exercise
Solve Problems 23, 27, 32, 34, 42 and 71 at the end of Chapter 7 in the
textbook (pages 189-192).
Answer to 32: 0.16 m, 0.94 m
Answer to 34: 4.5 103 J
Answer to 42: 33.0, 0.808 m/s

Assignment 2 Questions
9. Three masses are positioned on a frictionless surface, as shown.
Initially, mass m1 (1.0 kg) moves with a velocity of 2.0 m/s to
the right, mass

m2 (2.0 kg) is at rest, and mass m3 (3.0 kg)

Page 193 of 261

moves to the left with a velocity of 0.50 m/s. First, mass

m1

m2 and recoils to the left.


Afterwards, mass m3 collides with mass m2 and sticks. Calculate,
collides elastically with mass
a. the speeds of masses

m1 and m2 after the first collision.

b. the speeds of masses

m2 and m3 after the second

collision.

10. A 2.0 kg ball moving with a speed of 3.0 m/s hits, elastically, an
identical stationary ball as shown. If the first ball moves away
with angle 30 to the original path, determine
a. the speed of the first ball after the collision.
b. the speed and direction of the second ball after the
collision.

Before starting the next lesson, test your comprehension of the material covered in
this unit by turning back to the objectives. Make certain that you can meet each
objective. If you have difficulties, do not hesitate to contact your tutor to discuss the
problem.

Footnote
1

Note that all references to the textbook are to Physics: Principles with
Applications, 6th ed., by Douglas C. Giancoli (Prentice Hall: Upper Saddle River, NJ:
2005).
Top

STUDY GUIDE

Page 194 of 261

Lesson 16
Rotational Kinematics

hen you were a kid, you probably enjoyed playing with a little toy
called a yo-yo. It basically consists of a piece of string wrapped around a
small cylinder. If you hold the end of the string in your hand and then
release the cylinder, it will accelerate downward, unwrapping the string.
However, you will notice that the cylinder does not behave like a freely
falling object. The downward acceleration of the cylinder will be smaller
than the gravitational acceleration g. Furthermore, you will also notice that
the cylinder is also spinning about its center with increasing speed. So, how
do we describe the motion of a yo-yo? To answer this question you need to
understand rotational dynamics. In this lesson, however, we will begin by
introducing some important concepts that involve the definition of the
center of mass and the kinematics of rotational motion.

Center of Mass
In studying the motion of various objects and bodies, we have been
treating them as point masses (without explicitly saying so). However, all
objects in real life are three-dimensional extended bodies with the total
mass distributed over the entire volume of the object. This may create a
problem in specifying the location of the object. In other words, what point
inside (or outside) the object, is used to mark its exact position? Another
added complication is that a three-dimensional object may rotate or spin
while in motion, causing different parts of the object to have different
velocities.
That, however, should not stop us from studying motion of objects as point
masses. For a given object, we can always find a point, called the center
of mass, where we can imagine that the total mass of the object is
concentrated. If the object rotates while in motion, the axis of rotation will
go through the center of mass. Motion of an extended body can then be
analyzed into a translational motion (of the center of mass) and a
rotational motion about the center of mass. In other words, what we have
been doing so fartreating extended bodies as point massesare not
meaningless exercises. We just need to add the extra complication due to
rotation.
When the mass of a body is evenly distributed over its entire volume, the

Page 195 of 261

center of mass coincides with the geometric center of the body. For
example, the center of mass of a uniform solid sphere is at the center of
the sphere. The center of mass of a hollow sphere of a uniform thickness
will also be at the center of the sphereoutside of the body. So the center
of mass of a body can be inside or outside the body.

Objectives
After completing this lesson, you should be able to
1. compute the center of mass of a many-body system.
2. define angular displacement, angular velocity and angular
acceleration.
3. explain the relationships among linear displacement, velocity and
acceleration, and their angular counterparts, when a body moves
in a circular path.
4. write out the kinematic equations of uniform rotational motion.
5. use the kinematic equations to describe and predict the angular
motion of a body rotating (or rolling) with constant acceleration.

Reading Assignment
Read section 7-8 in the textbook.1

Angular Quantities
Opening a door is an activity we perform countless times a day without
consciously thinking about it. However, as you will see in this lesson and
the next, the simple operation of opening a door has far-reaching
consequences in physics.
Consider a door attached to a hinge as shown in the diagram above. To
open the door, you must apply a force to push or pull the door. As you
apply the force, the door swings open about the hinge. The motion of the
door as a whole and the motions of parts of the doorthe door knob, a nail
hammered into the door, atoms and molecules that make up the door or
the nailmove in arcs of different radii with a common center at the hinge.
Consider two points A and B on the door as shown. In a time interval t, as
the door swings open, A will move to A. Similarly, B will move to B. Since

Page 196 of 261

point B moves
a larger
distance than
point A during
the same time
interval, the
linear velocity
of point B is
larger than
that of point A.
Thus, as the
door swings
open, different
parts of the
door move in
circular paths
of different
radii with
different linear
velocities.
If we tried to
analyze the motion of the door using our knowledge of linear motion, we
would end up with a horrendously complex set of equations that would
frighten even the most accomplished mathematician. This does not mean
that we cannot tackle such problems. Physicists always claim that physics
is simple, and if we are to believe them the motion of rigid bodies should
not be complicated. The clue to simplicity is realizing that each point of a
rigid body executing rotational motion about a fixed axis moves in a
circular path of constant radius. If we can find a simple way of analyzing
circular motionas we did for linear motionwe can hope that rotational
motion of rigid bodies will be less complex than it appears at first glance.
You must have noted, when you were studying kinematics of linear motion,
that the characteristics of a motionspeed, velocity, acceleration, kinetic
energy, etc.are not directly measurable quantities. One calculates
velocity, for example, from the measurement of displacement over a time
interval. Description of motion thus depends on how we choose to measure
the displacement and the time interval.
Let us look again at the hinged door shown above. The linear displacement
of point A during the time interval t is l which is the length of the arc
AA. The average speed of point A is then equal to

Page 197 of 261

Note that we have mentioned the speed, not the velocity, of point A. This is
because in a circular motion, the direction of the velocity is always tangent
to the circle. Even when the speed remains constant, the direction of the
velocity changes from instant to instant. So, the direction of the velocity at
A is different from that at A. This characteristic holds for motion along any
curved path.
Life becomes somewhat simpler in such cases if we choose to work with
angular displacement instead of linear displacement. In the example
above, the angular displacement of point A during the time interval t is
the angle . Actually, since the door is a rigid body, all points inside the
door perform the same angular displacement during the same time
interval. The relationship between the angular displacement and the
linear displacement of point A, for example, is given by

The unit of angular displacement is radians (denoted by rad). As the body


completes one revolution, it travels a linear distance equal to the
circumference of the circle, that is, 2r. The corresponding angular distance

is then (2r)/r = 2 rad. There are two other common units for angular
displacement: degrees (denoted by ), and revolutions (denoted by rev).
These different units are related as follows:
1 rev = 2 rad = 360,
1 rad = (360/2) = 1/(2) rev,
1 = 1/360 rev = (2/360) rad.
The average angular velocity,
defined as

of the door during the time interval t is

For sufficiently small t, the definition above can be used for the
instantaneous velocity. The units of angular velocity are obviously rad/s,
/s or rev/s. Angular and linear velocities are related as follows:

If the angular velocity of the door changes by an amount equal to


during the time interval t, then we have angular acceleration, denoted by
, defined as the rate of change of angular velocity. Assuming constant

Page 198 of 261

acceleration, we can write,

where the appropriate units are rad/s2, /s2 or rev/s2. The relation
between angular and linear acceleration is given by

With these definitions of angular displacement, velocity and acceleration,


we can develop the kinematics and dynamics of a body moving in a circular
path in terms of angular quantitiesjust as we developed the kinematics
and dynamics of linear motion in the previous lessons.
To develop a system to describe rotational motion, we proceed as we did in
Lesson 3 when we developed the kinematics of linear motion. We will
consider only uniform rotational motion, that is, motion along a circular
path with constant angular acceleration . Assume the initial angular
position (at t = 0) of the rotating object is 0 and the initial angular velocity
is

0. If the angular velocity after time t is , then we have the following

kinematic equations of uniform rotational motion:

The above equations allow us to describe and predict the motion of a body
moving along a circular path with constant acceleration. Note their
similarity to the equations for linear motion with constant acceleration.

Reading Assignment
Read section 5-4 in the textbook. Then, read sections 8-1, 8-2 and 8-3.

Chapter 7: problem 48 (Page 191)

Top

The CM of an empty 1050-kg car is 2.50 m behind the front of the car. How
far from the front of the car will the CM be when two people sit in the front
seat 2.80 m from the front of the car and three people sit in the back seat
3.90 m from the front? Assume that each person has a mass of 70.0 kg.
Solution

Page 199 of 261

In this problem we assume that the centers of mass for the car and for
each passenger are located on the same horizontal line. This makes the
problem one-dimensional. Conveniently, we set the origin of the x-axis at
the front of the car. With this choice of coordinates, the car is located at
xcar = 2.50 m, the front-seat passengers are located at xf = 2.50 m and the
back-seat passengers are located at xb = 2.50 m. With two people sitting in
the front and three in the back (see the diagram above), the center of
mass of the car with the passengers is located at

where

mp = 70.0 kg is the mass of each passenger. So, the center of mass

of the car, when the passengers get in, shifts towards the back by 24 cm.
End of Solution

Chapter 7: problem 49 (Page 191)


A square uniform raft, 18 m by 18 m, of mass 6800 kg, is used as a ferry
boat. If three cars, each of mass 1200 kg, occupy its NE, SE and SW
corners, determine the CM of the loaded ferry boat.
Solution

Page 200 of 261

In this two-dimensional problem, we will set the origin of the x-y coordinate
system at the geometric center of the raft, which coincides (due to
uniformity) with its center of mass. Cars A, B and C are located (see the
diagram below) at the NE, SE and SW corners, respectively. Since the raft is
square with dimensions 18 m 18 m, the locations are as follows:
Raft Car A
Car B
Car C
Location x = 0 x = 9.0 m x = 9.0 m x = 9.0 m
r
A
B
C
yr = 0 yA = 9.0 m yB = 9.0 m yC = 9.0 m

Since the mass of the raft is


1200 kg, the x and
three cars are

mr = 6800 kg and each car has a mass m =

y coordinates of the center of mass of the raft with the

Page 201 of 261

and

End of Solution

Chapter 8: problem 4 (Page 219)


The blades in a blender rotate at a rate of 6500 rpm. When the motor is
turned off during operation, the blades slow to rest in 3.0 s. What is the
angular acceleration as the blades slow down?
Solution
When the motor is turned off, the angular speed of the blades decreases
from an initial value 0 = 6500 rev/min to a final value = 0 in a time
interval t = 3.0 s. The change in the angular speed of the blades is,
therefore, equal to

This information allows us to calculate the average angular acceleration of


the blades as follows:

Notice that if the acceleration of the blades is assumed uniform, then we


could substitute into the kinematic equation for rotational motion: = 0
+

t. The calculated acceleration will be the same as above.

Page 202 of 261

End of Solution

Chapter 8: problem 10 (Page 219)


Due to the earths rotation, what is the linear speed of a point
a. on the equator,
b. on the Arctic Circle (latitude 66.5 N), and
c. at a latitude of 45.0 N, due to the Earths rotation?
Solution
We know that the Earth spins about its axis, making a complete rotation in
one day. In other words, a point fixed to the Earth performs an angular
displacement = 2 in a time interval t = 24 h. Since the rotation is
uniform, the angular speed of any point on the surface (or within) the Earth
is equal to

The linear distance covered by a point on the Earths surface in one day is
equal to the circumference of a circle whose radius is the perpendicular
distance between the point and the axis of rotation. In the diagram below,
we notice that points at lower latitudes move in larger circles than points at
higher latitudes. Therefore, the linear speed of a point on the equator is
much larger than the linear speed of a point close to one of the poles.
a. A point on the equator rotates in a circle whose radius is equal to
the radius of the Earth (i.e., r = 6.38106 m). Therefore, the
linear speed of point a (see the diagram above) is equal to

b. Points on the Arctic Circle rotate in a (smaller) circle of radius

Page 203 of 261

rb =
rcos66.5
= 2.54
106m,
with a
linear
speed

vb = rb
= 185
m/s.

c. A point
located at
a latitude
of 45.0 N
rotates in
a circle of
radius
rc =
rcos45 =
4.51
106m,
with a linear speed

vc = rc = 328 m/s.
End of Solution

Chapter 8: problem 12 (Page 219)


A 70-cm-diameter wheel accelerates
uniformly about its center from 130 rpm to 280 rpm in 4.0 s.
a. Determine its angular acceleration.
b. Determine the radial and tangential components of the linear
acceleration of a point on the edge of the wheel 2.0 s after it has
started accelerating.
Solution

Page 204 of 261

Top

a. In this problem the wheel


accelerates uniformly from an
initial speed

0 = 130 rev/min = 13.61


rad/s
to a final speed

= 280 rev/min = 29.32


rad/s
in a time period t = 4.0 s.
By substituting these values
into the kinematic equation
= 0 + t, we get
29.32 rad/s = 13.61 rad/s +

4.0 s,

and the angular acceleration of the wheel is calculated to be

b. Notice that the wheel in this problem does not perform a uniform
circular motion. This is because in addition to the change in
direction, the magnitude of the velocity of any point on the wheel
continues to increase with time. After 2.0 s of uniform
acceleration, the wheel reaches an angular speed

For a point on the edge, this corresponds to a linear speed

Such a point can turn and move in a circular arc of radius r at this
instant in time, if it has a centripetal (or radial) acceleration equal
to

Page 205 of 261

The change in speed (or magnitude of the velocity) is caused by


the tangential acceleration, which is parallel to the direction of
the velocity. For a point on the edge, this acceleration will be
tangent to the wheel and is equal to

Notice that for a particular point on the wheel, the radial


acceleration increases with time due to the increase in the
rotation speed, while the tangential acceleration remains
constant.
End of Solution

Chapter 8: problem 18 (Page 219)


A wheel 33 cm in diameter accelerates uniformly from 240 rpm to 360 rpm
in 6.5 s. How far will a point on the edge of the wheel have traveled in this
time?
Solution
At time t = 0, the initial angular speed of the wheel is 240 revolutions per
minute. After accelerating for 6.5 s, the wheel reaches a final angular
speed of 360 revolutions per minute. To calculate the linear distance
traveled, we need to know the number of revolutions made by the wheel.
This is because each point on the edge travels a linear distance equal to the
circumference of the wheel in each complete revolution.
To begin, we summarize the information we have about the motion of the
wheel as follows:

Page 206 of 261

As an intermediate step, the angular acceleration of the wheel is calculated


by substituting into the kinematic equation for angular motion, = 0 + t
such that
360 rev/min = 240 rev/min +

0.1083 min,

and

If we now substitute into the kinematic equation = 0 +


get the angular displacement

0t + t2, we

This result indicates that every point on the wheel makes 32.5 revolutions
during the acceleration period. Since the wheel has a radius r = 0.165 m,
the distance traveled by a point on the edge is equal to

End of Solution

Chapter 8: problem 20 (Page 219)


A small, rubber wheel is used to drive a large pottery wheel, and they are
mounted so that their circular edges touch. The small wheel has a radius of
2.0 cm and accelerates at the rate of 7.2 rad/s2; it is in contact with the
pottery wheel (radius 25.0 cm) without slipping. Calculate
a. the angular acceleration of the pottery wheel.
b. the time it takes the pottery wheel to reach its required speed of
65 rpm.

Page 207 of 261

Solution

a. Let us focus on the point where the wheels touch (see the
diagram above). At this point, the edges of the two wheels move
together with the same linear speed in the same direction. If the
linear speed of one wheel increases, the linear speed of the other
wheel increases at the same rate. Therefore the tangential
acceleration (artan) of a point on the edge of the rubber wheel is
equal to the tangential acceleration (aptan) of a point on the edge
of the pottery wheel, or

artan = aptan.
Since tangential and angular accelerations are related by

a = r,

the above equation becomes


Rrr = Rpp,
where, Rr and Rp are the radii of the rubber and pottery wheels
respectively. The angular acceleration of the pottery wheel is,
therefore, equal to

Page 208 of 261

b. The time it takes the angular velocity of the pottery wheel to


increase from the initial value 0 = 0 to the final value = 65
rpm = 6.807 rad/s, is calculated by substituting into the
kinematic equation = 0 + t, such that
6.807 rad/s = 0 + 0.576 rad/s2 t,
and

Notice that the unrounded value of the acceleration is used in the


above calculation.
End of Solution

Exercise
Solve Problems 46 and 50 at the end of Chapter 7 in the textbook
(page 191). Then, solve Problems 5, 7, 14, 15 and 19 at the end of
Chapter 8 in the textbook (page 219).
Answer to ch7-46: 0.44 m
Answer to ch7-50: 3.8 l0 from the left edge of the smallest cube
Answer to ch8-14: (a) 1.5 104rad/s2, (b) 6.2 104m/s2, 8.1
103m/s2

Assignment 2 Questions
11. The angular speed of a wheel increases at a constant rate of 1.5
rad/s2. During a certain 3.2 s time interval, it makes four
complete revolutions. If the wheel started from rest, how long
has it been in motion at the start of this time interval?

Page 209 of 261

12. An airplane makes a circular turn of radius 9.0 km at a constant


speed of 650 km/h. Calculate the magnitudes of the planes
a. angular velocity
b. centripetal acceleration
c. angular acceleration
d. tangential acceleration
Before starting the next lesson, test your comprehension of the material covered in
this unit by turning back to the objectives. Make certain that you can meet each
objective. If you have difficulties, do not hesitate to contact your tutor to discuss the
problem.

Footnote
1

Note that all references to the textbook are to Physics: Principles with
Applications, 6th ed., by Douglas C. Giancoli (Prentice Hall: Upper Saddle River, NJ:
2005).
Top

STUDY GUIDE

Lesson 17
Rotational Dynamics

hat causes rotational motion? We have one obvious answerwe


need a force, as we saw in the example of opening a door. If we do not
apply any force on a door, it will stay shut forever. The application of a
force generates an angular acceleration about the axis of rotationthe
hingesand the door starts picking up angular velocity and executes a
rotational motion. It seems that we are back to Newtons Second Law.

Not really! In case of rotational motion, angular acceleration is not


determined by the applied force alone. The point on the solid body at which
the force is applied is also a determining factor. You must have noticed that
opening the door by pushing at a point close to the hinges takes a

Page 210 of 261

superhuman effortso we always push a door at the edge, far from the
hinges. The distance between the axis of rotation and the point at which
the force is applied is also a critical factor in the resulting rotational motion.
Let us analyze this situation further.
Suppose we apply a force F, as
shown in the diagram to the
right, at the edge of the door.
The direction of F, called the
line of action of F, goes through the hinge. The door will not movethe
hinge exerts an equal and opposite force on the door. We can safely
conclude that if the line of action of the applied force passes through the
axis of rotation, it does not produce a rotational motion.
What will happen if the force F is
applied at an angle , as shown
in the diagram? If we resolve F
into the two components F cos
and F sin , it is clear that the
component F cos will have no
effect on the rotational motion of
the door since its line of action
will pass through the hinge. Only
F sin will be effective in
producing rotational motion of
the door. Our experience tells us that a larger force will open the door
quicker, i.e., the angular acceleration will be directly proportional to F sin
. The angular acceleration will also be directly proportional to r, the
distance between the point of application of F sin and the axis of rotation,
the hinge. So, we can write

r F sin ,
where the symbol means directly proportional to.
The quantity r F sin , which can also be written as F r sin , is called
the torque of the force F about the axis of rotation O. From the diagram
above, we see that the quantity r sin , referred to as the lever arm, is
equal to the perpendicular distance between the line of action of the force
and the axis of rotation. We use the Greek letter (pronounced tau) to
denote torque, and define it as follows: the torque due to a force of
magnitude F about a point O is the product of F and the lever arm.
Equivalently, we can also define the torque as the product of the radius
arm r and the component of F perpendicular to r. In SI units, torque is
measured in N m.

Page 211 of 261

So we have , i.e., the angular acceleration of a rotating body is


proportional to the torque. This relationship is similar to Newtons second
law, where F, for linear motion. Our next task is to search for a
relationship in rotational motion that will be similar to F =

ma.

Consider a mass m moving in a


circular path of radius r, with the
center at O. Suppose that a force
F is applied on the mass,
perpendicular to r, as shown in
the diagram to the right.
According to Newtons Second
Law, the mass experiences a
linear acceleration tangent to the
circular path, and is related to F
by F = ma. The mass also
experiences a torque about O,
generated by the applied force, with a value equal to = r F. Since the
linear acceleration is related to the angular acceleration by a = r, we can
write

The quantity mr2) is called the moment of inertia of the mass m about
the point O. Writing I for the moment of inertia, we get the equation we
have been looking for:

= I.
According to this equation, when = 0, then = 0. In other words, in the
absence of a torque a body either will not have a rotational motion or will
have a uniform rotational motion with constant angular velocity. A net
torque will produce an angular acceleration , determined by the moment
of inertia I of the body about the axis of rotation, with value = /I. Once
the angular acceleration has been determined, the subsequent rotational
motion of the body due to the applied torque can be described using the
kinematic equations of rotational motion. The similarity between = I and
F = ma needs some further exploration. The torque plays the same role in
rotational motion that F plays in linear motion; angular acceleration
replaces linear acceleration a, and the moment of inertia I is the equivalent
of the mass m. There is, however, one complication. The mass m in linear

Page 212 of 261

motion is a simple constant quantity, while the moment of inertia is not so


simple. For the same mass, the moment of inertia of a body will have
different values depending on the location of the axis of rotation and the
shape of the body.
To see how the moment of
inertia of a rigid body is
computed, let us look at a simple
examplea thin loop, of radius R
and mass M, rotating about an
axis through its center at O with
an angular acceleration . We
can imagine that the loop is
divided into very small sections
with masses m1, m2, . . . etc.,
such that the mass M of the disk
is equal to M = m1 + m2 + . . .
etc. Each tiny section can then
be treated as a small particle rotating about O. Since the loop is a rigid
body, then every part of the loop is rotating with the same angular
acceleration and is located at the same distance R from the axis of
rotation at the center. Therefore, the torque acting on m1 is 1 = I1 =

m1R2, and that acting on m2 is 2 = I2 = m2R2, and so on. The total


torque on the rotating disk is then

where
I = MR2
is the moment of inertia of the thin loop about its center O.
It is important to note that the moment of inertia of the same loop about
another axis of rotation will not be the same as above. This is because the
distances between the various parts of the loop and the axis of rotation will
be different when the position of O changes. A little bit of calculus, which is
beyond the scope of this course, helps us to compute the moments of
inertia of solid bodies of regular shapes from the above definition. For
example, the moment of inertia of a solid uniform disk of mass M and

Page 213 of 261

radius R about its center is I = MR2. If we move the axis of rotation, the
same disk will have a different moment of inertia. The moment of inertia, in
SI units, is measured in kg m2.

Objectives

Top

After completing this lesson, you should be able to


1. define torque and explain the meaning of lever arm.
2. define moment of inertia.
3. explain the dynamics of rotational motion.

Reading Assignment
Read sections 8-4, 8-5 and 8-6 in the textbook.1

Chapter 8: problem 23 (Page 220)


A person exerts a force of 55 N on the end of a door 74 cm wide. What is
the magnitude of the torque if the force is exerted
a. perpendicular to the door.
b. at a 45 angle to the face of the door?
Solution
Torque is equal to the magnitude
of the applied force times the
lever armthe perpendicular
distance between the direction of
the force and the axis of rotation.
a. When the force F = 55 N is
applied perpendicular to the
door, the lever arm is equal
to the doors width, or r =
0.75 m. The magnitude of
the torque due to the force
is then

Page 214 of 261

b. When the force is applied at an angle of 45 (see the diagram


above) the lever arm is shorter (i.e., r sin ) and the magnitude
of the torque reduces to

End of Solution

Chapter 8: problem 26 (Page 220)


The bolts on the cylinder head of an engine require tightening to a torque
of 88 m N. If a wrench is 28 cm long, what force perpendicular to the
wrench must the mechanic exert at its end? If the six-sided bolt head is
15 mm in diameter, estimate the force applied near each of the six points
by a socket wrench (Fig. 8-41 in the textbook).

Solution: In this problem a force Fw is applied at the end of a wrench,


whose axis of rotation is at the other end. Since the applied force is
perpendicular to the wrench, its length (rw = 0.28 m) is equal to the lever
arm, and the magnitude of the generated torque is given by the equation
= rw Fw. To generate a torque of 88 m N, the mechanic must apply a
force equal to

Page 215 of 261

The torque generated by the applied force is transferred to the bolt, which
feels six forces of equal magnitude (Fb) corresponding to the six points of
contact. As shown in the diagram above, these forces are tangent to a
circle of radius rb = 7.5 mm centered about the axis of rotation, and they
generate a torque

= 6 (rb Fb). Since = 88 m N, then the force

applied near each of the six points by the socket wrench is equal to

Notice that the force felt by the bolt is 37 times greater than the force
applied at the end of the wrench. However, both generate the same torque
due to the differences in arm lengths.
End of Solution

Chapter 8: problem 30 (Page 220)

Top

A potter is shaping a bowl on a potter's wheel that is rotating at constant


angular speed (Fig. 8-42 in the textbook). The friction force between her
hands and the clay is 1.5 N total.
a. How large is her torque on the wheel, if the diameter of the bowl
is 12 cm?
b. How long would it take for the potter's wheel to stop if the only
torque acting on it is due to the potter's hand? The initial angular
velocity of the wheel is 1.6 rev/s, and the moment of inertia of
the wheel and the bowl is 0.11 kg m2.
Solution
a. Since the force of friction due to the potters hands is tangent to
the walls of the circular bowl, the lever arm is just the radius r =
6.0cm. If the direction of the bowls rotation is taken to be
positive, then the torque generated by the friction force is
negative. This is because it attempts to cause a rotation in the
opposite direction. The applied torque, due to friction, is therefore
equal to

Page 216 of 261

b. If this is the only torque acting on the system, then the angular
velocity is not constant. According to Newtons Second Law for
Rotation, the rotating wheel and bowl experience angular
acceleration () such that such that = I, where I is the
moment of the rotating system. Since both and I are known, we
can calculate as follows

If the initial angular speed is equal to

0 = 1.6 rev/s = 10.05 rad/s,


then the time it takes the wheel to stop ( = 0) is calculated by
substituting into the kinematic equation = 0 + t, as follows
0 = 10.05 rad/s + (0.818rad/s2) t.
Solving for t, we find that the wheel stops after

t = 12s.
End of Solution

Chapter 8: problem 34 (Page 221)


A grinding wheel is a uniform cylinder with a radius of 8.50 cm and a mass
of 0.580 kg. Calculate
a. its moment of inertia about its center.
b. the applied torque needed to accelerate it from rest to 1500 rpm
in 5.00 s if it is known to slow down from 1500 rpm to rest in
55.0 s.
Solution
a. According to Figure 8-21 (in the textbook) the moment of inertia
for a uniform cylinder of radius R and mass M is equal to
.
So, the moment of inertia of the grinding wheel is equal to

Page 217 of 261

b. To increase the angular speed of the grinding wheel from rest (0


= 0) to = 1500 rpm = 157.1 rad/s in 5.00 s, its angular
acceleration must be equal to

To generate such acceleration, according to Newtons Second


Law, the net torque acting on the wheel must be equal to,

The additional information in the problem indicates that the


applied torque (applied) is not the only torque acting on the
wheel. The forces of kinetic friction generate a torque (friction)
that acts to resist the rotation of the wheel, thus causing it to
slow down when the applied force is removed. The magnitude of
the frictional torque is constant and vanishes only when the
wheel stops rotating. The net torque acting on the wheel is
therefore equal to

net = applied + friction.


In the absence of any applied torque, the forces of friction act
alone and cause the wheel to slow down from an initial angular
speed 0 = 1500 rpm = 157.1 rad/s to a final speed = 0, in
55.0 s. Since this corresponds to an angular acceleration

Page 218 of 261

the frictional torque is calculated to be

Now, we can calculate the applied torque on the wheel as follows:

End of Solution

Chapter 8: problem 40 (Page 221)


A helicopter rotor blade can be considered a long, thin rod, as shown in
Fig. 8-46 (in the textbook).
a. If each of the three rotor helicopter blades is 3.75 m long and has
a mass of 160 kg, calculate the moment of inertia of the three
rotor blades about the axis of rotation.
b. How much torque must the motor apply to bring the blades up to
a speed of 5.0 rev/s in 8.0 s?
Solution
a. According to Figure 8-12 (in the textbook) the moment of inertia
for a long uniform rod of length L and mass M rotating about its
end is equal to
.
So, the total moment of inertia of the three helicopter rotor
blades is equal to

b. To start the blades from rest (0 = 0) and bring their angular

Page 219 of 261

Top

speed to = 5.0 rev/s = 31.42 rad/s in 8.0 s, the angular


acceleration must be equal to

which corresponds to a torque

applied by the motor.


End of Solution

Chapter 8: problem 41 (Page 221)


An Atwood's machine consists of two masses,

m1 and m2, which are

connected by a massless inelastic cord that passes over a pulley (Fig. 8-47
in the textbook). If the pulley has radius R and moment of inertia I about
its axle, determine the acceleration of the masses m1 and m2, and compare
to the situation in which the moment of inertia of the pulley is ignored.
Solution
Let us begin with the simple case
of a massless and a frictionless pulley. In other words, we will ignore the
presence of the pulley except for the purpose of transferring the motion
from one mass to the other through a massless cord. In this case, no
torque is required to rotate the pulley, and the tension (FT) is the same
along the cord.
Intuitively, we know that if one of the masses is greater than the other, the
system will accelerate towards the heavier mass. The cord, in this case, will
have just the right tension such that the two masses move with the same
speed and have the same acceleration. From Newtons Second Law, we
write
FT

m1g = m1a,

Page 220 of 261

and

m2g FT = m2a.
By adding the two equations
above, we get

m2g m1g = m1a + m2a,


which gives to the following
expression for the acceleration
.
Notice that if the masses are
equal the acceleration becomes
equal to zero and no motion will
occur. Also, if we substitute m1 = 0, in the expression above, we get

a = g.

This means that if one of the masses is removed, the other one will
perform a free fall.
If the mass of the pulley cannot be ignored, then its rotation will affect the
motion of the other two masses and must be taken into consideration. In
particular, the pulleys rim has the same linear acceleration as the other
two masses. This means that the pulley experiences angular acceleration
equal to = a/r. The difference in the cords tension on both sides of the
pulley produces the required torque (FT2R FT1R) for such angular
acceleration. By substituting into Newtons Second Law for Rotation, we
have
(FT2 FT1) R = I,
or
.
The corresponding equations for the other two masses are
FT1

m1g = m1a

and

m2g FT2 = m2a.


By adding the two equation above, we get

Page 221 of 261

FT1 FT2 +

m2g m1g = m1a + m2a,

or
FT1 FT2 = (m1 +

m2)a (m2 m1)g.

Comparing this equation with the one corresponding to the pulley, we


notice that both are equal, or

By rearranging terms the equation above becomes

and we get the following expression for the acceleration

Notice that when the mass of the pulley is ignored, its moment of inertia
becomes zero, and the above expression reduces to that of a massless
pulley.
End of Solution

Exercise
Solve Problems 24, 27, 29, 32, 37 and 78 at the end of Chapter 8 in the
textbook (pages 220-224).
Answer to 24: 1.4 m N clockwise
Answer to 32: 1.2 1010m
Answer to 78:
(a)
(b)

Page 222 of 261

Assignment 2 Questions
13. A square plate has a side length of 1.2 m. An applied torque of
300 N m gives the plate an angular acceleration of 4.8 rad/s2
about an axis through the centre of the plate. Determine the
plates mass.
14. A light string is wrapped around a solid cylinder and a 300 g
mass hangs from the free end of the string, as shown. When
released, the mass falls a distance 54 cm in 3.0 s.
a. Draw free-body diagrams for the block and the cylinder.
b. Calculate the tension in the string.
c. Calculate the mass of the cylinder.

Before starting the next lesson, test your comprehension of the material covered in
this unit by turning back to the objectives. Make certain that you can meet each
objective. If you have difficulties, do not hesitate to contact your tutor to discuss the
problem.

Footnote
1

Note that all references to the textbook are to Physics: Principles with
Applications, 6th ed., by Douglas C. Giancoli (Prentice Hall: Upper Saddle River, NJ:
2005).
Top

STUDY GUIDE

Lesson 18

Page 223 of 261

Conservation Laws in Rotational Motion

t is very convenient to separate the general motion of a threedimensional object into translational and rotational parts. This allows us to
make use of the laws and equations developed for linear (or translational)
motion with few additions and modifications.

Objectives
After completing this lesson, you should be able to
1. discuss the Law of Conservation of Mechanical Energy as applied
to a body having both translational and rotational motion.
2. define angular momentum.
3. discuss the Law of Conservation of Angular Momentum.

Rotational Kinetic Energy


When a rigid body rotates about a fixed axis, each particle in the body
executes a circular motion, with the same angular velocity about the axis.
The linear velocities, however, will be different for different particles,
depending on their distances from the axis of rotation. Let us return to the
example (from the previous lesson) of the thin loop of mass M and radius R
rotating about its center O with an angular velocity . As before, we will
imagine that the disk is made up of small masses m1, m2, . . . etc., at a

m1 + m2 + . . . . The linear velocity of each


section of the loop will be v = R, with the direction along the tangent of

distance R from O, with M =

the loop. Summing up the kinetic energies of all loop sections, we find the
rotational kinetic energy of the rotating loop to be

Actually, the equation above is a general result and applies to any rigid
body, even though it was derived for the case of a thin loop. Note that the

Page 224 of 261

unit of measurement of rotational kinetic energy is kg m2 s2 = J, as we


would expect.

The definition of kinetic energy can now be extended to include the kinetic
energy of rotational motion. Consider a rigid body, of mass, M, that is
moving and rotating at the same time (e.g., a rolling wheel or a falling yoyo). The total kinetic energy of such a body is equal to the sum of the
translational and rotational kinetic energies. More precisely, we have

where vCM is the linear velocity of the center of mass, ICM is the moment of
inertia about an axis through the center of mass and is the angular
velocity of rotation about this axis.
The Law of Conservation of Mechanical Energy can also be extended to
include the kinetic energy of rotational motion. For example, consider a
wheel rolling down an inclined slope, as shown in the diagram above.
Assume that the wheels linear and angular velocities at points A and B are
v1, 1 and v22, respectively. According to this law, the total mechanical
energy of the wheel at point A must be equal to the total mechanical
energy at point B. Mathematically, this is written as

where I is the moment of inertia of the wheel of mass M about its center.
The two quantities h1 and h2 are the heights of the center of the wheel at
points A and B, respectively.

Angular Momentum

Page 225 of 261

Top

The angular momentum L of a mass


p, where

m about an axis is defined by L = r

p = mv is the linear momentum of the mass and r (see

diagram below) is the perpendicular distance of the axis of rotation from


the direction of the momentum p. With this definition, we can derive the
following relation for the angular momentum:

Notice that that the


component of the velocity v
cos , which is directed away
from the axis, does not
generate any angular
displacement. Therefore, the
angular displacement
depends only on v sin the
component of the velocity
perpendicular to r. The
relation above is actually a
general one and can be
applied to any rotating rigid
body. The unit of angular
momentum in SI system is kg m2 s1.
The torque can be expressed in terms of the change in angular momentum
as follows:

This equation is similar to F = p/t, Newtons second law, and states that
the rate of change of angular momentum is equal to the net torque on the
body when = 0 and L/t = 0, which means that L does not change with
time. This is the Law of Conservation of Angular Momentumif the net
torque on a body is zero, then the bodys total angular momentum remains
constant.

Reading Assignment

Page 226 of 261

Read sections 8-7 and 8-8 in the textbook.1

Chapter 8: problem 48 (Page 222)


A sphere of radius 20.0 cm and mass 1.80 kg starts from rest and rolls
without slipping down a 30.0 incline that is 10.0 m long.
a. Calculate its translational and rotational speeds when it reaches
the bottom.
b. What is the ratio of translational to rotational KE at the bottom?
(Avoid putting in numbers until the end so you can answer.)
c. do your answers in a. and b. depend on the radius of the sphere
or its mass?
Solution
Let us begin with the simpler
case of a sphere sliding down
a frictionless incline without
rolling. According to the Law of
Conservation of Mechanical
Energy, the initial gravitational
potential energy of the sphere
is converted into
(translational) kinetic energy
at the bottom of the incline,
such that

Therefore, the velocity of the sphere when it reaches the bottom of the
incline is given by
.
a. The mechanical energy of the sphere is also conserved if it rolls
down the incline without sliding. However, in this case only part
of the initial gravitational potential energy is converted into
translational kinetic energy. The remaining part changes into
rotational kinetic energy of the sphere about its center. So, we
write
PE = KEtranslational + KErotational

Page 227 of 261

or
,
where v and are the translational and the angular speeds of the
sphere at the bottom of the incline, respectively. Substituting

(moment of inertia of a solid sphere) and


equation above we get

= v/r into the

.
By canceling the mass of the sphere and then solving for v, we
get the following expression for the translational velocity of a
rolling sphere at the bottom of the incline
.
Notice that it is smaller than the speed derived above for a sliding
sphere with no rotation. Also, notice that v is independent of both
m and r. This means that any solid sphere released at the top of
the incline reaches the bottom with same translational speed,
regardless of the mass and size of the sphere.
To calculate the angular speed, we divide the expression by the
radius as follows

.
So,

depends on the size of the sphere but not on its mass.

By substituting h = 10.0 m sin 30.0 = 5.00 m, and r = 0.200


m in the two equations above, we get the following numerical
values for the translational and the angular velocities at the
bottom of the incline

v = 8.37 m/s
and

= 41.8 rad/s.
b. The ratio of KEtranslational to KErotational we get

Page 228 of 261

So, the translational kinetic energy of a rolling solid sphere is


always 2.5 times larger than its rotational kinetic energy,
regardless of the mass and radius.
c. The answer to this part is included in parts a. and b. above.
End of Solution

Chapter 8: problem 49 (Page 222)


Two masses,

Top

m1 = 18.0 kg and m2 = 26.5 kg, are connected by a rope that

hangs over a pulley (as in Fig. 8-47 in the textbook). The pulley is a
uniform cylinder of radius 0.260 m and mass 7.50 kg. Initially, m1 is on the

m2 rests 3.00 m above the ground. If the system is now


released, use conservation of energy to determine the speed of m2 just
ground and

before it strikes the ground. Assume the pulley is frictionless.


Solution
In problem 41 (see previous lesson), we derived an expression for the
acceleration of a similar system by analyzing the forces acting on each
object and applying Newtons Second Law. Since the pulley is a uniform
cylinder of radius r and mass mp, the derived expression takes the form

.
By substituting for the masses, we get the following value for the
magnitude of the acceleration of each object in the system

a = 1.726 m/s2.
So, when the system is released,

m2 starts from rest (v0 = 0) and falls to

Page 229 of 261

the ground (y y0 = 3.00 m) with acceleration a = 1.726 m/s2. Notice


that m2 has a negative downward acceleration. This information is sufficient
to calculate the final speed of

m2, just before it strikes the ground. By

substituting into the kinematic equation v2 = v02 + 2a(y y0), we get

v2 = 0 + 2 (1.726 m/s2) (3.0 m),


and

v is calculated to be

v = 3.22 m/s.
Sounds good. However, is it possible to solve this problem using the
Principle of Conservation of Mechanical Energy? Yes. This is what we are
going to do next. Since the pulley is frictionless, then the total mechanical
energy of the system is conserved. In other words, the potential energy
plus the kinetic energy at the beginning of the motion is equal to the
potential energy plus the kinetic energy at any later time.

Since the system starts moving from rest, its initial kinetic energy is equal
to zero. Also, since the mass m1 is at the ground level when the system is
released, the initial potential energy is equal to the initial gravitational
potential energy of the mass m2. So, the initial mechanical energy of the
system is equal to
Ei =

m2gh,

Page 230 of 261

where

h = 3.00m is the initial height of the mass m2 above the ground

level.
When the system is released, it accelerates towards the heavier mass, and
every moving object acquires kinetic energy. Also, when the lighter mass
(m1) rises, it acquires gravitational potential energy. Therefore, final
mechanical energy of the system, just before the mass

m2 strikes the

ground, is given by

where we have substituted I = m r2 for the moment of inertia of the


p
uniform cylindrical pulley, and = v/r for its angular speed. Since Ei = Ef,
we write
,
or
.
By solving for the final speed, we get the following expression:

and the following numerical value

which is in agreement with the result calculated above using a different


method. Notice the convenience of solving this problem using conservation
of mechanical energy.
End of Solution

Chapter 8: problem 50 (Page 222)


A 2.30-m-long pole is balanced vertically on its tip. It starts to fall and its
lower end does not slip. What will be the speed of the upper end of the pole

Page 231 of 261

just before it hits the ground? [Hint: Use conservation of energy.]


Solution
Since the pole falls down from
rest, its initial kinetic energy is
equal to zero. The gravitational
potential energy of the pole is
equal to mgh, where h is the
height of its center of mass (CM)
above the ground. Since the CM
is in the middle of the pole, its
initial potential energy is mg
L/2. Therefore, the initial total
mechanical energy of the pole,
before it starts to fall, is equal to
Ei = mgL.
While falling down, the pole
rotates about its lower end, thus lowering the CM and increasing the
angular speed (see the diagram). When the pole is about to hit the ground,
its final gravitational potential energy becomes equal to zero, thus
converting all its initial potential energy into rotational kinetic energy. So,
the final mechanical energy of the pole is given by
Ef = I2.
From Figure 8-21 in the textbook, we see that the moment of inertia of a
long uniform rod of length L and mass m is equal to
I = mL2.
So, the final mechanical energy of the pole becomes
.
By equating the initial and the final energies (conservation of mechanical
energy), we get the equation
.
The poles mass cancels in this equation, and we get the following
expression for the angular speed of the pole just before it hits the ground

Page 232 of 261

The upper end of the pole rotates in a circle of radius L about the lower
end. So, the linear speed of the upper end, just before the pole hits the
ground, is equal to

End of Solution

Chapter 8: problem 52 (Page 222)


a. What is the angular momentum of a 2.8-kg uniform cylindrical
grinding wheel of radius 18 cm when rotating at 1500 rpm?
b. How much torque is required to stop it in 6.0 s?
Solution:
a. The angular momentum (L) of a rotating object is equal to its
moment of inertia times the angular velocity, or
L = I.
The rotation speed of the grinding wheel is given in terms of
revolutions per minute. However, to have uniformity of units in
the calculations, it is more appropriate to express in terms of
radians per second. So, we write

Since the grinding wheel is a solid uniform cylinder of radius r


and mass m its moment of inertia is given by
I = mr2.
By substituting back into the first equation, the angular
momentum of the grinding wheel is calculated to be

Page 233 of 261

Top

b. The torque applied on an object is equal to the rate of change of


its angular momentum, or
.
When the grinding wheel stops rotating, its angular momentum
changes by

If this occurs in t = 6.0 s, then the applied torque on the wheel


must be equal to

The negative torque indicates that it acts in the opposite direction


of the wheels rotation.
End of Solution

Chapter 8: problem 53 (Page 222)


A person stands, hands at his side, on a platform that is rotating at a rate
of 1.30 rev/s. If he raises his arms to a horizontal position (Fig. 8-48 in the
textbook), the speed of rotation decreases to 0.80 rev/s.
a. Why?
b. By what factor has his moment of inertia changed?
Solution
a. When a person raises his arms, he extends them away from his
body, which increases the distance between the center of mass of
each arm and the axis of rotation. As a result, the moment of
inertia of the person increases. However, due to the absence of

Page 234 of 261

external torque, the angular momentum of the person is


conserved. In other words, the moment of inertia times the
angular speed (or I) remains constant. So, we argue then that
the angular speed must decrease as the person raises his arms.
b. We will use subscript i to denote the values of the relevant
quantities before the person raises his arms and subscript f for
the values of the same quantities when he has raised his arms.
From conservation of angular momentum, we get
Iii = If

or

So, the moment of inertia of the person increases by


approximately 60% when he raises his arms to a horizontal
position.
End of Solution

Chapter 8: problem 60 (Page 222)


A uniform disk turns at 2.4 rev/s around a frictionless spindle. A nonrotating rod of the same mass as the disk and length equal to the disk's
diameter is dropped onto the freely spinning disk (Fig. 8-49 in the
textbook). They then both turn around the spindle with their centers
superposed. What is the angular frequency in rev/s of the combination?
Solution
Before the rod is dropped, the initial angular momentum (Li) of the system

is equal to that of the rotating disk only. This is because the angular speed
of the rod, just before it comes in contact with the disk, is zero. So, we
write
Li = Idi,

i is the initial angular speed of the rotating disk and Id is its


moment of inertia. If f is the final angular speed of the disk-rod
where

combination, then the final angular momentum is expressed as


Lf = Idf + Irf ,

Page 235 of 261

Ir is the moment of inertia of the rod about its center.


Since no external torque is
involved, the total angular
momentum of the disk and the
rod is conserved. This is because
after coming in contact with the
rotating disk the rod picks up
angular momentum and causes
the disk to lose angular
momentum by exactly the same
amount. So, Li and Lf are equal,

and we write

Idi = Idf + Irf .


In this problem, the length of the rod is equal to the disks diameter, and
both objects have the same mass (m). Referring to Figure 8-21 (page 208)
in the textbook, the moments of inertia for both objects are
Id = mr2
and
.
Substituting back into the angular momentum equation, we get
,
or
.
So, the final angular speed (or angular frequency) of the combination is
equal to

End of Solution

Exercise

Page 236 of 261

Solve Problems 45, 47, 51, 56 and 61 at the end of Chapter 8 in the
textbook (pages 221-222).
Answer to 56: 1.4 rev/s

Assignment 2 Questions
15. A solid sphere of mass 6.0 kg is mounted on a vertical axis and
can rotate freely without friction. A massless cord is wrapped
around the middle of the sphere and passes over a 1.0 kg pulley
and is attached to block of mass 4.0 kg, as shown. What is the
speed of the block after it has fallen 80 cm? Treat the pulley as
solid cylinder.

16. Consider a uniform rod of mass 12 kg and length l.0 m. At its


end, the rod is attached to a fixed, friction-free pivot. Initially the
rod is balanced vertically above the pivot and begins to fall (from
rest) as shown in the diagram. Determine,
a. the angular acceleration of the rod as it passes through
the horizontal at B.
b. the angular speed of the rod as it passes through the
vertical at C.

Before starting the next lesson, test your comprehension of the material covered in
this unit by turning back to the objectives. Make certain that you can meet each

Page 237 of 261

objective. If you have difficulties, do not hesitate to contact your tutor to discuss the
problem.

Footnote
1

Note that all references to the textbook are to Physics: Principles with
Applications, 6th ed., by Douglas C. Giancoli (Prentice Hall: Upper Saddle River, NJ:
2005).
Top

STUDY GUIDE

Lesson 19
Static Equilibrium

conomists talk about equilibrium between supply and demand;


stockbrokers discuss market equilibrium; physicists lecture on
thermodynamic equilibrium. Whatever the context, we associate the word
equilibrium with balance and stability. We imagine situations where
objects do not move around, although we expect them to. The Empire
State Building stays at the same location for the same reason a picture
stays suspended on your living room wall. Absence of motion seems to be
an ideal picture of equilibrium.
The study of equilibrium is of substantial interest to physicists. Insight
acquired in this area through the laws of physics has provided the
foundations for a variety of disciplinesfrom engineering to social sciences
to medicine. The level of this course will not allow a thorough study of
equilibrium. We will confine ourselves to the basic analysis of objects in
static equilibrium and applications of such analyses in real-life situations.

Objectives
After completing this lesson, you should be able to

Page 238 of 261

1. define static equilibrium.


2. describe the conditions of equilibrium.
3. use the conditions of equilibrium to solve problems involving
static equilibrium.

Equilibrium
The Latin word equilibrium means equality of forces. In the simple situation
in which a body is subjected to two forces, equality of forces will obviously
mean that the two forces are equal and opposite, so that the net force on
the body equals zero. When there are three forces on a body and the sum
of any two is equal and opposite to the third, we can say that there is an
equality of forces. If n forces are acting on a body, then equality of forces
will mean that sum of (n 1) forces must be equal and opposite to the
remaining one. In other words, an object is in equilibrium when the vector
sum of the forces acting on it is zero, or F = 0. This requirement is known
as the first condition of equilibrium. If all objects in this universe were
point masses, this condition is the only one that would be needed to ensure
that a body was in equilibrium.
Most objects in real life,
however, are extended bodies,
and the first condition alone
cannot guarantee the equilibrium
of a body. To see this clearly,
consider a plank balanced at its
center on a pivot. If two equal
and opposite forces of magnitude
F are applied at the two edges of
the plank, the net force on the
plank will be zero. So, the first
condition of equilibrium is
satisfied, and the plank will not have any translational motion. However,
from our experience we know that the plank will rotate in the counterclockwise direction, as shown in the diagram. To ensure absence of motion,
we need a second requirement: that the net torque on a body must be
equal to zero. This requirement, = 0, is known as the second condition
of equilibrium.
We will confine ourselves to analyzing objects in static equilibrium, with no
translational (v = 0) or rotational ( = 0) motions. Furthermore, we will
consider only those situations in which all the forces lie in a common plane,
which we will assume to be the x-y plane. With these restrictions, we will

Page 239 of 261

need to deal with only three scalar quantities: the x- and y-components of
the forces, and the torque of the forces about any point in the x-y plane.
The two conditions of equilibrium thus provide us with the following three
equations:

which must be satisfied simultaneously. Note that the axis of torque


equation is arbitrary and can be chosen anywhere.

Reading Assignment
Read sections 9-1, 9-2 and 9-3 in the textbook.1

Stability and Balance

Top

Suppose an extended body is in static equilibrium with v = 0 and = 0.


How will it react if it is slightly displaced from its equilibrium position? To
answer this question, we need to know the position of the center of gravity
of the body relative to the point of support. When the center of gravity lies
below the point of support, the body invariably returns to its original
position of equilibrium, and we say that the body is in stable equilibrium. If
the center of gravity lies above the point of support, then the body moves
away from its original equilibrium position, and the state of equilibrium is
called unstable equilibrium. The body is said to be in neutral equilibrium if
it continues to stay in its new position, which will be the case if the center
of gravity coincides with the point of support.

Reading Assignment
Read section 9-4 in the textbook.

Chapter 9: problem 2 (Page 247)


Calculate the torque about the front support post (B) of a diving board (
Fig. 9-42 in the textbook), which is exerted by a 58-kg person 3.0 m from
that post.
Solution

Page 240 of 261

While standing at the end of the


board, the diver exerts a
downward force equal to her
weight, as shown in the diagram
to the right. To calculate the
torque about the front support
(B), we simply multiply divers
weight (mg) by the distance between the diver and point B (or lever arm),
such that

End of Solution

Chapter 9: problem 12 (Page 248)


Find the tension in the two wires supporting the traffic light shown in
Fig. 9-46 (textbook).
Solution
From the diagram to the right,
we see three forces acting on the
traffic light: its weight and the
tensions (T1 and T2) in the
supporting wires. Since the
traffic light is under static
equilibrium, the net force acting
on it is equal to zero. At the
components level, this means
that the summation of all the
horizontal components of the
three forces is equal to zero (or
Fx = 0), and the summation of
all the vertical components is
also equal to zero (or Fy = 0).
So, we write

Fx = T1cos 37 T2cos 53 = 0
and

Page 241 of 261

Fy = T1sin 37 + T2sin 53

mg = 0

These are two equations in two unknowns that we will solve simultaneously
to find the tensions in the wires. A simple method is to solve the first
equation for T2 (in terms of T1) such that

and then substitute into the second equation as follows


.
This equation is easily solved, because it has one variable (T1). We get

Substituting back into T2, we get

So, the tension in the two wires supporting the traffic light (rounded to two
significant figures) are 190 N in the wire making an angle 37 with the
horizontal, and 260 N in the wire making an angle 53 with the horizontal.
End of Solution

Chapter 9: problem 18 (Page 248)


Calculate
a. the tension FT in the wire that supports the 27-kg beam shown in
Fig. 9-52 (in the textbook).
b. the force FW exerted by the wall on the beam (give magnitude
and direction).
Solution
At the beginning, note that there
are three unknowns in this problem: the tension in the wire (FT) and the
magnitude (FW) and direction () of the force exerted by the wall. So,

Page 242 of 261

mathematically, three
independent equations are
required to find unique values of
these unknowns and, therefore,
solve the problem. The
conditions of static equilibrium
provide such equations, where
the net horizontal force, the net
vertical force and the net torque
acting on the beam must all be
equal to zero.
From the free-body diagram
above, we see that stability against translational motion gives
Fx = FWcos FTcos 40 = 0,
and
Fy = FWsin + FTsin40

mg = 0.

Note that the two equations above contain all three unknowns. Before
writing the third equation ( = 0), we must select an axis of rotation.
However, since the location of such axis is arbitrary, it is more convenient
to select it such that the calculated net torque contains the minimum
number of unknowns. In our problem, a suitable location is at the end of
the beam supported by the wall. With this selection, FW does not contribute
to the net torque on the beam since the lever arm of this force is equal to
zero. The resulting torque equation is, therefore, given by

a. Note that the length of the beam (L) cancels out in the equation
above, which allows us to solve for FT such that

After rounding to two significant figures, the tension in the wire is


equal to 210 N.
b. Substituting this result into the equations for the horizontal and
vertical components of the net force, we get

Page 243 of 261

and

which are the

x and y components of the force (FW) exerted by

the wall on the end of the beam. Taking the ratio of these two
equations we get

which gives a value for equals to


.
The magnitude FW is calculated by substituting in (for example)
the equation of horizontal net force as follows:

We see in this problem that in order for the beam to be in static


equilibrium, both FT and FW must have the same magnitude
(210 N) and must make an angle 40 with the beam such that the
horizontal components of both forces are directed in opposite
directions towards the center of the beam.
End of Solution

Chapter 9: problem 24 (Page 249)

Top

The two trees in Fig. 9-58 are 7.6 m apart. A backpacker is trying to lift his
pack out of the reach of bears. Calculate the magnitude of the force F that

Page 244 of 261

he must exert downward to hold a 19-kg backpack so that the rope sags at
its midpoint by (a) 1.5 m and (b) 0.15 m.
Solution
If we ignore friction between
the rope and the tree branch,
then the force F exerted by
the backpacker is equal to
the tension in the rope. Since
the backpack is in
equilibrium, then all the
forces acting on it must add
up to zero. In particular (see the free-body diagram to the right), the
vertical components of the rope tension must balance the weight of the
backpack. So, we write
F sin + F sin =

mg

where is the angle that the rope makes with the horizontal. Solving the
equation above for F, we get

a. If the rope sags at its midpoint by 1.5 m, then the angle is


given by

and the tension in the rope is equal to

b. If the rope sags at its midpoint by 0.15 m, then

and the tension is the rope is equal to

End of Solution

Page 245 of 261

Chapter 9: problem 26 (Page 249)


A uniform ladder of mass m and length l leans at an angle against a
frictionless wall (Fig. 9-60 in the textbook). If the coefficient of static
friction between the ladder and the ground is , determine a formula for
the minimum angle at which the ladder will not slip.
Solution
Before we begin the solution, it is important to analyze and understand the
different forces acting on the ladder. In the free-body diagram below, we
see the downward force of gravity (mg) acting at ladders center of mass,
or the midpoint in this case. This is balanced by the upward normal force
(FN) acting on the lower end of the ladder. The other force acting at the
lower end is the horizontal force of static friction, which resists any slipping
motion on the ground. Since the wall is frictionless, it will not resist any
vertical motion of ladders upper end. So, there is only one reaction force
(FW) perpendicular to the wall, as shown in the diagram.
These four forces must cancel each other, resulting in a zero net force, to
prevent any translational motion. Also, the net torque on the ladder must
be equal to zero to prevent any rotational motion. So, we write

Notice that in this problem, it is more convenient to calculate the net


torque about the lower end of the ladder. In this case, FN and Ffr have zero
lever arm and do not contribute to the net torque. However, the location of
the rotation axis is arbitrary and can be chosen anywhere without affecting
the final answer.
Let us begin by the torque equation above and solve for the reaction force
(FW) generated by the wall, such that

If you substitute different values of in this equation, you will notice that
FW increases as decreases. Since Ffr is equal to FW according to the first

equation above, then the force of static friction also increases as the ladder
leans at smaller angles, and we write

Page 246 of 261

However, the maximum force of static friction generated between the


ground and the ladder is equal to
max Ffr =

FN = mg

where we have used FN = mg from the second equation above. Therefore,


to prevent the ladder from slipping, FW should at most be equal to this
maximum value, or

In this equation,

mg cancels out and we get

So, the minimum angle at which the ladder will not slip is equal to

Note that the result is independent of the mass and length of the ladder.

Page 247 of 261

End of Solution

Chapter 9: problem 28
(Page 250)
A person wants to push a lamp
(mass 7.2 kg) across the floor,
for which the coefficient of
friction is 0.20. Calculate the
maximum height x above the
floor at which the person can
push the lamp so that it slides
rather than tips (Fig. 9-62 in the
textbook).
Solution: We begin by drawing a
free-body diagram as shown
below. The lamp, of course, does
not move in the vertical
direction. So, the net vertical
force acting on it is equal to zero,
or
Fy = FN

mg = 0

From this equation, we see that the normal force applied on the lamp by
the floor is equal to
FN =

mg

Consider the situation where the lamp stands still and is not being pushed
(that is FP = 0). In this case, the upward normal force acts at the center of
the base along the same line of the weight, and the net torque on the lamp
is equal to zero. Assume now that the person applies a horizontal force (FP)
on the lamp at a small height x just above the floor. When this happens,
the normal force shifts slightly away from the center towards the edge of
the base, to balance the added torque. When the applied force is increased
until the lamp is on the verge of tipping, the normal force will act at the
edge of the base as shown in the diagram to the right. In this case, the net
torque on the lamp, calculated about the edge of the base, is equal to
=

mg 0.12m FP x = 0

From this equation, we see that the maximum horizontal force applied on
the lamp at a height x above the floor is given by

Page 248 of 261

A greater force at this height causes the lamp to rotate about the edge and
tip.
In the relation above we see that as x increases, the maximum applied
force decreases. However, FP cannot be decreased indefinitely since it
should at least be equal to the force of kinetic friction (Ffr =
maintain the sliding motion. Therefore, the equality

mg) to

is satisfied at the maximum height above the floor at which the person can
push the lamp and slide it along the floor without causing it to tip. Solving
for this maximum height, we get

Notice that we did not need the mass of the lamp in the calculation.
End of Solution

Exercise
Solve Problems 1, 7, 9, 11, 23, 27 and 31 at the end of Chapter 9 in the
textbook (pages 247-250).

Assignment 2 Questions
Note: These are the last questions of Assignment 2. After completing this assignment
please submit it to your tutor for marking.

17. A uniform vertical beam of mass 40 kg is acted on by a horizontal


force of 520 N at its top and is held, in the vertical position, by a
cable as shown.
a. Draw a free-body diagram for the beam, clearly labeling
all of the forces acting on it.
b. Calculate the tension in the cable?
c. Determine the reaction forces acting on the beam by

Page 249 of 261

the ground?

18. A 30 kg neon sign is suspended by two cables, as shown. Three


neighborhood cats (5.0 kg each) find the sign a comfortable
place. Calculate the tension in each cable when the cats are in
the positions shown.

Before starting the next lesson, test your comprehension of the material covered in
this unit by turning back to the objectives. Make certain that you can meet each
objective. If you have difficulties, do not hesitate to contact your tutor to discuss the
problem.

Footnote
1

Note that all references to the textbook are to Physics: Principles with
Applications, 6th ed., by Douglas C. Giancoli (Prentice Hall: Upper Saddle River, NJ:
2005).
Top

STUDY GUIDE

Page 250 of 261

Lesson 20
Review II

y completing Lesson 19, you have covered all the material of this
course and should start preparing yourself to write the final examination.
This lesson contains a review of the main concepts covered in the second
half of the course and suggests a number of practice problems that you can
do as a self test. Please consult your Course Manual for information about
the final examination. You can also find a sample exam with a formula
sheet and solutions as well as additional practice problems under the
Resources section of the course.

Work and Kinetic Energy


The work done on an object by a constant force is given by
W=F

d cos ,

where F is the magnitude


of the applied force, d is
the magnitude of the
displacement and is the
angle between the direction
of the force and the
direction of the
displacement. The diagram
to the right shows two
situations in which a box is
displaced on a horizontal surface by applying forces in different directions.
In the first situation, the force and the displacement are in the same
direction, i.e., = 0. In the second situation, the applied force makes an
angle > 0 with direction of the displacement. Notice that if the angle
between the force and the displacement is = 90, then no work is done
on the object. In the diagram above, the weight of the box is perpendicular
to the direction of its displacement. So, the work done by the force of
gravity is zero. There are situations when the applied force is in the
opposite direction of the displacement, i.e., = 180. An example is the
force of friction between the box and the surface in the diagram above. The
work done by friction is negative in this case. From the definition of work,

Page 251 of 261

we can see that it has the unit (kg m2 s2 = N m = J), called the Joule
An object of mass

m moving with a velocity v has kinetic energy equal to

KE = mv2
Work has the effect of changing the kinetic energy. So, the final kinetic
energy of an object is equal to its initial kinetic energy plus the net work
done on the object, such that
KEf = KEi + W.
Notice that no work is done by the centripetal force on an object moving in
uniform circular motion. This is because the displacement of the object and
the centripetal force are always perpendicular to each other. So, there is no
change in the kinetic energy, and the speed stays constant.

Gravitational Potential Energy


The work done against gravity in raising an object by a vertical distance
above the ground is given by

This work is stored in the form of potential energy, and we say that the
potential energy of the object increases by the amount
W = PEf PEi =

mgh

This quantity is also equal to the work done by the force of gravity on the
object when it is released and allowed to fall down a distance h towards the
ground. Actually, the relation above is more general and can be used to
calculate the work done by gravity on an object moving between two
vertical heights, regardless of the path. The work done by gravity is the
same in the four situations shown below. The force of gravity is an example
of conservative forces whose work depends only on the initial and final
positions and the path followed by the object. Therefore, the work done by
such forces on object that moved in a closed path (and returned to its
initial position) is zero. Nonconservative forces, such as the force of kinetic
friction, do not have this property, and no potential energy is associated
with them.

Page 252 of 261

Conservation of Mechanical Energy

Top

The total mechanical energy (E) of an object at a particular time is defined


as the summation of the kinetic and potential energies of that object. In
the absence of nonconservative (also called dissipative) forces, the total
mechanical energy is conserved and does not change with time or location.
This is called the law of conservation of mechanical energy, and it is
written as
E = KE + PE = constant.
So, if the only force acting on object is the force of gravity, then we can
write initial mechanical energy = final mechanical energy,

mgh0 + mv02 = mgh + mv2.


The work done by nonconservative forces is equal to the change in the tota
mechanical energy of the object, i.e.,

Conservation of Linear Momentum


The impulse given to a certain object is equal to the product of the average
force (F) and the time interval (t) during which the force acts:

Page 253 of 261

Impulse = F t.
The linear momentum (p) of an object is equal to the product of the
objects mass (m) and its velocity (v):
P=

mv.

Both impulse and linear momentum are vector quantities and have the
units N s = kg m/s. These two quantities are related by the impulsemomentum theorem, which states that impulse is equal to the change in
the linear momentum.
If the net external force acting on a system of objects is equal to zero, then
the impulse given to the system is also equal to zero. As a result, the total
linear momentum of the system is conserved. The internal forces (that the
objects within the system exert on each other) change the momenta of the
individual objects. However, the total vector sum of the momenta is a
constant.

Collisions
Two colliding objects form a system. If this system is isolated (i.e., no
external forces), then the total linear momentum is conserved, and we can
write
momentum before collision (p) = momentum after collision (p).
However, the total kinetic energy of the system is not always conserved,
and collisions are classified into two types: elastic collisions, where the
kinetic energy of the system is conserved, and inelastic collisions, where
the kinetic energy of the system is not conserved. For collisions in two
dimensions, the total linear momentum of the system is conserved in both
the x- and the y-directions, separately.

Center of Mass
The center of mass (CM) is a
point representing the average location of the mass of a particular system.
For example, consider the three particles shown in the diagram below,
which are distributed in a two-dimensional space. Intuitively, we expect the
center of mass to be located somewhere in the triangle joining the three
particles. We also expect it to be closer to the heavier particle (m2 in this
example). To locate the center of mass more precisely, we calculate the x-

Page 254 of 261

and the y-components of its


position as follows:

Notice that if the total momentum of a system of particles is conserved,


then the velocity of the center of mass is constant.

Linear and Angular Quantities

Top

Consider an object moving in a circular path, as shown in the diagram


below. In the case of uniform circular motion, the object will move with a
constant speed equal to v = l/t, where t is the time it takes the particle
to move from point A to point B on the circular path. The velocity vector,
on the other hand, is not constant because the object continuously changes
its direction of motion. The acceleration, therefore, is perpendicular to the
direction of the velocity and always points towards the center of the circular
path. This type of acceleration is called centripetal acceleration and its
magnitude is equal to

aR = v2/r.
If the rotation of the object is not uniform, then the magnitude of its
velocity may change with time, in addition to the continuous change in
direction. Since the object may have different speeds at points A and B, the
quantity v = l/t will represent the average linear speed between these
two points. In addition to the centripetal acceleration, which is responsible
for changing the direction of the velocity, there is another type, called
tangential acceleration, which is responsible for changing the magnitude of

Page 255 of 261

the velocity of the rotating


object. If the change in the
objects speed between points A
an B is v, then the tangential
(or linear) acceleration of the
object is given by atan = v/t.
Notice that in the case of uniform
circular motion

atan = 0.
The angular displacement () of
a rotating object is equal to the
difference between its initial and final angular positions. In the diagram
above, this is equal to = 0, and is related to the linear
displacement by the equation

Angular velocity is defined as the rate of change of angular position with


time. In the diagram above, the average angular velocity of the rotating
object between points A and B is given by

For a sufficiently small time interval, we can derive the following relation
between linear and angular velocities:

Angular acceleration is defined as the change of angular velocity with time,


and it is related to linear acceleration as follows:

The units of angular displacement are degree (), radian (rad) or number
of revolutions (rev), which are related as follows:
1 rev = 2 rad = 360.
The units commonly used for angular velocity are rad/s and rev/s. For
angular acceleration we use rad/s2 or rev/s2. As a matter of convention,
angular displacement, angular velocity and angular acceleration in the

Page 256 of 261

counter-clockwise direction are assigned positive values and vice versa.


Notice that the centripetal acceleration of a rotating object is related to its
angular velocity as follows:

Equations of Rotational Kinematics


There is a clear analogy between the variables and equations of linear and
rotational kinematics as shown in the tables below.
Initial
Final
Initial Final
Constant
Time
displacement displacement velocity velocity acceleration interval
Linear

x0

v0

Rotational

Linear Kinematic
v = v0 + at

Rotational kinematics
= 0 + t

x = (v0 + v)t

= (0 +

x = x0 + v0t + at2

= 0 +

)t

0t + t2

v2 = v02 + 2a(x x0) 2 = 02 + 2a( 0)

Rolling Motion
Rolling combines both translational and rotational motions. Consider a
wheel of radius r rolling with constant speed on a horizontal surface, as
shown in the diagram below. If the wheel covers a distance d in a time t,
then its translational velocity is equal to v = d/t. The rotational motion is
not independent from the translational motion in this case. As seen from
the diagram, the wheel rotates a circular distance l = d during the same
time interval. Since the tangential velocity of a point on the rim is equal to
vtan = l/t, then we can see that both the translational and the tangential
velocities are equal. However, vtan = r, where

Page 257 of 261

is the angular velocity of

the wheel. So, for a rolling object we have

v = vtan = r.
So, each point on the rim
performs two types of motions
simultaneously. It moves with
the wheel in the horizontal
direction, with a translational
velocity v, and at the same time
it rotates about the center of the
wheel with a tangential velocity,
also equal to v. The net velocity
of the point is then equal to the vector sum of both types of velocities, as
shown in the diagram below. Notice that the point at the bottom of the
wheel has a net velocity equal to zero, while that on the top moves with a
net velocity equal to v.
Similarly, if the wheel accelerates, we can show that the linear acceleration
of its translational motion is related to the angular acceleration of its
rotational motion through the equation

a = r.

Rotational Dynamics

Top

The tendency of an object to rotate depends on the applied force and on


the perpendicular distance (called the lever arm) between the force and the
axis of rotation. This leads to the introduction of a new quantity, torque,
defined as
Torque () = applied force (F) lever arm (l).

Page 258 of 261

The torque is assigned a positive value if it tends to produce a rotation in


the counter-clockwise direction, and is assigned a negative value if it tends
to produce a rotation in the clockwise direction.
Rotational dynamics is described by the angular version of Newtons
Second Law, which states that the net torque on an object is proportional
to its angular acceleration such that

= I,
where I is the moment of inertia of the rotating object about the specified
axis of rotation.
If a body rotates (by an angle ) when a torque () is applied, then a work
(W) is done on the body, given by
W = .
The rotational kinetic energy of a body rotating with an angular velocity
is equal to

KE = I2.
In the absence of any external forces or torques, the total mechanical
energy of body, including rotational kinetic energy, is conserved and is
equal to
E = mv2 + I2 +

mgh,

where v is the translational speed of the center of mass, and


of the object relative to an arbitrary height.

h is the height

The angular momentum (L) of a body rotating about a fixed axis is the
product of the bodys moment of inertia and the angular velocity, such that
L = I.
In the absence of external torques, the total angular momentum of a
rotating body is conserved. Here is a comparison between equations of
linear and rotational dynamics:
Linear
Newtons Second Law F = ma

Rotational
= I

Work

W = Fd

W =

Kinetic energy

KE = mv2 KE = I2

Momentum

p = mv

L = I

Page 259 of 261

Static Equilibrium
A rigid body is in equilibrium if it has zero linear acceleration and zero
angular acceleration. The condition (a = 0) implies that the sum of all
external forces acting on the body is equal to zero, i.e.,
Fx = 0,
Fy = 0.
The other requirement ( = 0) implies that the sum of all external torques
about any axis is equal to zero, i.e.
= 0.

Self Test
Below are 21 problems selected from the textbook and grouped into three
sets. To test yourself, consider each set as a practice exam. Give yourself
limited time (approximately two and a half hours) and do not refer to any
material except the formula sheet while working on each set. It is
important to note that the purpose of the exam is to test your
understanding of the concepts introduced in the course. The actual exam
questions may be different from the problems below.
Problem Set 1 Chapter
Chapter
Chapter
Chapter
Chapter
Chapter
Problem Set 2 Chapter
Chapter
Chapter
Chapter
Chapter
Chapter
Problem Set 3 Chapter
Chapter
Chapter

2: problem 80
4: problem 43
6: problems 54 and 84
7: problem 79
8: problem 71
9: problem 5
3: problem 64
4: problem 19
6: problem 73
7: problems 8 and 70
8: problem 55
9: problem 13
2: problem 57
5: problem 35
6: problem 50

Page 260 of 261

Chapter 7: problem 66
Chapter 8: problems 55 and 86
Chapter 9: problem 20
Answer to Ch 2-80: 24.1 s, 230 m/s
Answer to Ch 3-64: 26.3 m/s, 3.8 m from the net, 0.714 s
Answer to Ch 6-50: (a) 16.0 m/s, (b) 1.06 N
Answer to Ch 6-54: 0.304
Answer to Ch 6-84: 184 m/s
Answer to Ch 7-8: 1.4 104 kg
Answer to Ch 7-66: 21 m
Answer to Ch 7-70: 5.47 m
Answer to Ch 8-86: (a) 1.61 rad/s2, (b) 23 s
Answer to Ch 9-20: 708 N, 580 N, 6 N
Good luck on your final examination!
Top

Page 261 of 261

You might also like