You are on page 1of 144

Basic Mathematics

Dimensional Analysis

R Horan & M Lavelle

The aim of this package is to provide a short


self assessment programme for students who
wish to learn how to use dimensional analysis
to investigate scientific equations.

Copyright c 2003 rhoran@plymouth.ac.uk , mlavelle@plymouth.ac.uk

Last Revision Date: June 2, 2004 Version 1.0


Table of Contents
1. Introduction
2. Checking Equations
3. Dimensionless Quantities
4. Final Quiz
Solutions to Exercises
Solutions to Quizzes

The full range of these packages and some instructions,


should they be required, can be obtained from our web
page Mathematics Support Materials.
Section 1: Introduction 3

1. Introduction
It is important to realise that it only makes sense to add the same sort
of quantities, e.g. area may be added to area but area may not
be added to temperature! These considerations lead to a powerful
method to analyse scientific equations called dimensional analysis.
One should note that while units are arbitrarily chosen (an alien
civilisation will not use seconds or weeks), dimensions represent
fundamental quantities such as time.
Basic dimensions are written as follows:
Dimension Symbol
Length L
Time T
Mass M
Temperature K
Electrical current I

See the package on Units for a review of SI units.


Section 1: Introduction 4

Example 1 An area can be expressed as a length times a length.


Therefore the dimensions of area are L × L = L2 . (A given area
could be expressed in the SI units of square metres, or indeed in any
appropriate units.) We sometimes write: [area]= L2
In some equations symbols appear which do not have any associated
dimension, e.g., in the formula for the area of a circle, πr2 , π is just
a number and does not have a dimension.
Exercise 1. Calculate the dimensions of the following quantities
(click on the green letters for the solutions).
(a) Volume (b) Speed
(c) Acceleration (d) Density

Quiz Pick out the units that have a different dimension to the other
three.
(a) kg m2 s−2 (b) g mm2 s−2
2 −2
(c) kg m s (d) mg cm2 s−2
Section 2: Checking Equations 5

2. Checking Equations
Example 2 Consider the equation
1
y = x + kx3
2
Since any terms which are added together or subtracted MUST have
the same dimensions, in this case y, x and 12 kx3 have to have the
same dimensions.
We say that such a scientific equation is dimensionally correct. (If
it is not true, the equation must be wrong.)
If in the above equation x and y were both lengths (dimension L) and
1/2 is a dimensionless number, then for the 12 kx3 term to have the
same dimension as the other two, we would need:
dimension of k × L3 = L
L
∴ dimension of k = 3 = L−2
L
So k would have dimensions of one over area, i.e., [k] = L−2 .
Section 2: Checking Equations 6

Quiz Hooke’s law states that the force, F , in a spring extended by a


length x is given by F = −kx. From Newton’s second law F = ma,
where m is the mass and a is the acceleration, calculate the dimension
of the spring constant k.
(a) M T−2 (b) M T2 (c) M L−2 T−2 (d) M L2 T2

Example 3 The expressions for kinetic energy E = 12 mv 2 (where


m is the mass of the body and v is its speed) and potential energy
E = mgh (where g is the acceleration due to gravity and h is the
height of the body) look very different but both describe energy. One
way to see this is to note that they have the same dimension.

Dimension of kinetic energy Dimension of potential energy


2 −1 2
1
2 mv ⇒ M (L T ) mgh ⇒ M (L T−2 ) L
= M L2 T−2 = M L2 T−2
Both expressions have the same dimensions, they can therefore be
added and subtracted from each other.
Section 2: Checking Equations 7

Exercise 2. Check that the dimensions of each side of the equations


below agree (click on the green letters for the solutions).
(a) The volume of a cylinder of (b) v = u + at for an object with
radius r and length h initial speed u, (constant)
V = πr2 h. acceleration a and final speed
v after a time t.
(c) E = mc2 where E is energy, (d) c = λν, where c is the speed of
m is mass and c is the speed light, λ is the wavelength and
of light. ν is the frequency

Note that dimensional analysis is a way of checking that equations


might be true. It does not prove that they are definitely correct.
Dimensional analysis would suggest that both Einstein’s equation E =
mc2 and the (incorrect) equation E = 12 mc2 might be true. On the
other hand dimensional analysis shows that E = mc3 makes no sense.
Section 3: Dimensionless Quantities 8

3. Dimensionless Quantities
Some quantities are said to be dimensionless. These are then pure
numbers which would be the same no matter what units are used (e.g.,
the mass of a proton is roughly 1850 times the mass of an electron no
matter how you measure mass).
Example 4 The ratio of one mass m1 to another mass m2 is dimen-
sionless:
m1 M
dimension of the fraction = =1
m2 M
The dimensions have canceled and the result is a number (which is
independent of the units, i.e., it would be the same whether the masses
were measured in kilograms or tonnes).
Note that angles are defined in terms of ratios of lengths. They
are therefore dimensionless! Functions of dimensionless variables are
themselves dimensionless.
Section 3: Dimensionless Quantities 9

Very many functions are dimensionless. The following quantities are


important cases of dimensionless quantities:

Trigonometric functions Logarithms


Exponentials Numbers, e.g., π

Note the following properties:


• Functions of dimensionless variables are dimensionless.
• Dimensionless functions must have dimensionless arguments.

Quiz If the number of radioactive atoms is found to be given as a


function of time t by
N (t) = N0 exp(−kt)
where N0 is the number of atoms at time t = 0, what is the dimension
of k?
(a) LT (b) log(T) (c) T (d) T−1
Section 3: Dimensionless Quantities 10

Exercise 3. Determine the dimensions of the expressions below (click


on the green letters for the solutions).
(a) In a Young’s slits experiment the angle θ of constructive interfer-
ence is related to the wavelength λ of the light, the spacing of the
slits d and the order number n by d sin(θ) = nλ. Show that this
is dimensionally correct.

(b) The Boltzmann distribution in thermodynamics involves the fac-


tor exp(−E/(kT )) where E represents energy, T is the tempera-
ture and k is Boltzmann’s constant. Find the dimensions of k.

Quiz Use dimensional analysis to see which of the following expres-


sions is allowed if P is a pressure, t is a time, m is a mass, r is a
distance, v is a velocity and T is a temperature.
P rt2
     2  
Pt Pr Pr
(a) log (b) log (c) log (d) log
mr m mt2 mtT
Section 4: Final Quiz 11

4. Final Quiz
Begin Quiz Choose the solutions from the options given.
1. Newton’s law of gravity states that the gravitational force be-
tween two masses, m1 and m2 , separated by a distance r is given
by F = Gm1 m2 /r2 . What are the dimensions of G?
(a) L3 M−1 T−2 (b) M2 L−2 (c) M L T−2 (d) M−1 L−3 T2
2. The coefficient of thermal expansion, α of a metal bar of length
` whose length expands by ∆` when its temperature increases by
∆T is given by ∆` = α`∆T. What are the dimensions of α?
(a) K−1 (b) L2 T−1 (c) L2 T−1 (d) L−2 K−1
3. The position of a mass at the end of a spring is found as a function
of time to be A sin(ωt). Select the dimensions of A and ω.
(a) L & T (b) L & Dimensionless
(c) sin(L) & T−1 (d) L & T−1

End Quiz Score: Correct


Solutions to Exercises 12

Solutions to Exercises
Exercise 1(a) A volume is given by multiplying three lengths to-
gether:
Dimension of volume = L×L×L
= L3
So [volume] = L3
(The SI units of volume are cubic metres.)
Click on the green square to return

Solutions to Exercises 13

Exercise 1(b) Speed is the rate of change of distance with respect


to time.
L
Dimensions of speed =
T
= L T−1
So [speed] = L T−1
(The SI units of speed are metres per second.)
Click on the green square to return

Solutions to Exercises 14

Exercise 1(c) Acceleration is the rate of change of speed with respect


to time
L T−1
Dimensions of acceleration =
T
= L T−2
So [acceleration] = L T−2
(The SI units of acceleration are metres per second squared.)
Click on the green square to return 
Solutions to Exercises 15

Exercise 1(d) Density is the mass per unit volume, so using the
dimension of volume we get:
M
Dimensions of volume =
L3
= M L−3
So [density] = M L−3
(The SI units of density are kg m−3 .)
Click on the green square to return 
Solutions to Exercises 16

Exercise 2(a) We want to check the dimensions of V = πr2 h. We


know that the dimensions of volume are [volume] = L3 . The right
hand side of the equation has dimensions:
dimensions of πr2 h = L2 × L = L3
So both sides have the dimensions of volume.
Click on the green square to return 
Solutions to Exercises 17

Exercise 2(b) We want to check the dimensions of the equation


v = u + at. Since v and u are both speeds, they have dimensions
L T−1 . Therefore we only need to verify that at has this dimension.
To see this consider:
[at] = (L T−2 ) × T = L T−2 × T = L T−1
So the equation is dimensionally correct, and all the terms have di-
mensions of speed.
Click on the green square to return

Solutions to Exercises 18

Exercise 2(c) We want to check the dimensions of the equation


E = mc2 . Since E is an energy it has dimensions M L2 T−2 . The right
hand side of the equation can also be seen to have this dimension, if
we recall that m is a mass and c is the speed of light (with dimension
L T−1 ). Therefore
[mc2 ] = M (L T−1 )2
= M L2 T−2
So the equation is dimensionally correct, and all terms that we add
have dimensions of energy.
Click on the green square to return 
Solutions to Exercises 19

Exercise 2(d) We want to check the dimensions of the equation


c = λν. Since c is the speed of light it has dimensions L T−1 . The right
hand side of the equation involves wavelength [λ] = L and frequency
[ν] = T−1 . We thus have
[λν] = L × T−1
which indeed also has dimensions of speed, so the equation is dimen-
sionally correct.
Click on the green square to return

Solutions to Exercises 20

Exercise 3(a) We want to check the dimensions of d sin(θ) = nλ.


Both d and λ have dimensions of length. The angle θ and sin(θ) as
well as the number n must all be dimensionless. Therefore we have
L×1 = 1×L
∴ [d sin(θ)] = [nλ]
So both sides have dimensions of length.
Click on the green square to return 
Solutions to Exercises 21

Exercise 3(b) The factor exp(−E/(kT )) is an exponential and so


must be dimensionless. Therefore its argument −E/(kT ) must also
be dimensionless.
The minus sign simply corresponds to multiplying by minus one
and is dimensionless. Energy, E has dimensions [E] = M L2 T−2
and temperature has dimensions K, so the dimensions of Boltzmann’s
constant are
M L2 T−2
[k] = = M L2 T−2 K−1
K
(So the SI units of Boltzmann’s constant are kg m2 s−2 o C−1 ).
Click on the green square to return

Solutions to Quizzes 22

Solutions to Quizzes
Solution to Quiz: kg2 m s−2 has
dimensions = M2 L T−2
It can be checked that all the other answers have dimension M L2 T−2 .
End Quiz
Solutions to Quizzes 23

Solution to Quiz: From Hooke’s law, F = −kx, we see that we can


write
F
k=
x
Now F = ma, so the dimensions of force are given by
[F] = M × (L T−2 )
= M L T−2
Therefore the spring constant has dimensions
M L T−2
[k] =
L
= M T−2
End Quiz
Solutions to Quizzes 24

Solution to Quiz: In N (t) = N0 exp(−kt), the exponential and its


argument must be dimensionless. Therefore kt has to be dimension-
less. Thus
dimensions of k × T = 1
So the dimension of k must be inverse time, i.e., [k] = T−1 .
End Quiz
Solutions to Quizzes 25

Solution to Quiz: First note that the argument of a logarithm must


be dimensionless. Now pressure is force over area, so it has dimensions
M L T−2
[P ] =
L2
= M L−1 T−2
2
Therefore the combination Pmrt
is dimensionless since
P rt2 (M L−1 T−2 ) × L × T2
 
=
m M
M
=
M
= 1
None of the other combinations are dimensionless and so it would be
completely meaningless to take their logarithms. End Quiz
North Berwick High School

Higher Physics
Department of Physics

Unit 1 Our Dynamic Universe

Section 1 Equations of Motion


Section 1 Equations of Motion

Note Making
Make a dictionary with the meanings of any new words.

Vectors and scalars

1. Describe the differences between vectors and scalars.


2. Make a table showing both including units. You must learn this.
3. Make a table showing the sign convention.
4. Describe a method of adding vectors (either scale diagrams or
trig). Don't forget top-to-tail. Include question 2 on page 22 as
an example.

Graphs
1. Draw the three sets of graphs on pages 4 - 6 of the booklet.
Make sure that you label them accordingly and that you
understand the differences between them.
2. Copy and answer question 5 on page 29 as an example.

Equations of Motion
1. Describe a method to measure acceleration using light gates.
2. Write down the three equations of motion.
3. Show the derivation of these equations.
4. Make a note of the techniques shown on page 15 and the
example on pages 16 and 17.
Section 1 Equations of Motion

Contents

Content Statements.........................................................................1
The physics of motion ...................................................................... 2
Vectors and scalars .......................................................................... 2
Scalar quantities .............................................................................. 2
Vector quantities ............................................................................. 3
Vectors and scalars: why do they matter? ........................................ 3
The sign convention ......................................................................... 3
Graphing motion and interpreting graphs......................................... 4
Interpreting displacement–time graphs ............................................ 7
Interpreting velocity–time graphs .................................................... 7
Investigating acceleration ................................................................ 8
The equations of motion .................................................................. 9
Direction matters ........................................................................... 13
Explaining good technique using an example .................................. 15
Example ......................................................................................... 16
Problems ....................................................................................... 18
Solutions ........................................................................................ 34
Content Statements
content Notes context

a)
Equations of motion Candidates should Light gates, motion sensors
for objects with undertake experiments to and software/hardware to
constant verify the measure displacement,
acceleration in a relationships shown in the velocity and acceleration.
straight line. equations. Using software to analyse
videos of motion.

b) Motion-time graphs Displacement-time graphs. Motion sensors (including


for motion with Gradient is velocity. wireless sensors) to enable
constant Velocity-time graphs. Area graphical representation of
acceleration. under graph is motion.
displacement. Investigate the variation of
Gradient is acceleration. acceleration on a slope
Acceleration-time graphs. with the
Restricted to zero and angle of the slope.
constant Motion of athletes and
acceleration. equipment used in sports.
Graphs for bouncing
objects and objects thrown
vertically
upwards.

c) Motion of objects Objects in freefall and the Investigate the initial


with constant speed movement of objects on acceleration of an object
or constant slopes should be projected
acceleration. investigated. vertically upwards (e.g.
popper toy)

1
Section 1 Equations of Motion

The physics of motion


Footballers, golfers, tennis players, runners, skiers – they all have
something in common. They have the ability to make split -second
decisions about how their actions will affect their performance: how
the curve of a ball will affect whether they score that crucial penalty,
whether a change of angle of the club will give them a hole-in-one.
They are making use of the physics of motion.
In the first section of ‘Our Dynamic Universe’ you will learn more
about the motion of objects: from raindrops to roller coasters. You will
be able to use the language of physics to describe and explain the
motion of bouncing balls and sky divers. You will understand the
principles of the physics of motion and explain it using wor ds,
diagrams, graphs and equations.

Vectors and scalars


Exploring the physics of motion, the terms acceleration, velocity and
displacement have been used.
What do these terms mean? Is displacement the same as distance? Is
velocity the same as speed?
Quantities can be defined as vectors or scalars. You will be familiar
with working with scalar quantities but you probably did not realise it.
You may also be familiar with some vector quantities and have
performed simple vector additions.

Scalar quantities
A scalar quantity is defined only by its magnitude (i.e. its size).
Examples of scalar quantities include:

Quantity Units
Distance m
Speed m s –1
Energy J
Time s
Mass kg

2
Vector quantities
A vector quantity is defined both by its magnitude (i.e. its size) and its
direction. Examples of vector quantities include:

Quantity Units
Displacement m
Velocity m s –1
Acceleration m s –2
Force N
Impulse Ns
Momentum kg m s –1
Weight N

Vectors and scalars: why do they matter?


In physics, we are studying the real, physical world and finding models
which explain our observations of motion in this world. Vectors give us
a method of describing motion in the real world.

So why does this matter? Consider a pilot coming in to land a n aircraft


in very windy conditions. As he tries to land the plane in the centre of
the runway, he must take into account the speed of the wind and its
direction, i.e. the velocity of the wind.

Consider the corrective action that will have to be taken by the pilot to
land an aircraft in a cross wind – and realise the importance of
understanding vectors!

The sign convention


Direction matters and so a system of defining direction is needed.
Normally, we use

upwards  positive
downwards  negative
right  positive
left  negative

3
Graphing motion and interpreting graphs
Although the ability to plot a graph accurately is important, it is
absolutely essential to be able to interpret graphs and to visualise the
shape of a graph for a given motion.

For an object that is stationary, the displacement–time, velocity–time


and acceleration–time graphs are shown below.

4
For an object that is moving with constant velocity, the
displacement–time, velocity–time and acceleration–time graphs are
shown below.

5
For an object that is accelerating uniformly, (i.e. moving with a
constant acceleration) the displacement–time, velocity–time and
acceleration–time graphs are shown below.

Each time you are presented with a graph, you should ask yourself two
key questions:

1. Is there any significance to the gradient of the graph?


2. Is there any significance to the area under the graph?

6
Interpreting displacement–time graphs
The gradient of the displacement–time graph is the velocity of the
object. Mathematically this is the equivalent of dividing the change in
the y-axis quantity by the change in the x-axis quantity: in this case
displacement divided by time. Displacement divided by time is rate of
change of displacement, which is velocity.

The area under the graph does not give us any meaningful information.
Mathematically this is the equivalent of multiplying the two quantities,
in this case displacement and time.

Interpreting velocity–time graphs


The gradient of the velocity–time graph is the acceleration of the
object: the steeper the line the greater the rate of change of velocity
(or acceleration) of the object.

The area under the graph is the displacement.

7
Investigating acceleration
Acceleration is defined as the rate of change of velocity per unit time.
From this definition, an equation can be written:

a = acceleration in metres per second per second


v u (m s –2 )
a
t v = final velocity in metres per second (m s –1 )
u = initial velocity in metres per second (m s –1 )
t = time in seconds (s)

The acceleration of an object can be investigated in a number of ways.


You should give consideration to the different methods, and the
advantages and disadvantages of each method for measurement of
acceleration in different circumstances. Methods of measurement of
acceleration include:

1. using light gates connected to a timing unit, with single -mask


and double-mask set-ups

2. using a motion sensor, which measures displacement with time


and from this calculates velocity and acceleration

3. using an accelerometer, which measures acceleration directly.

You should be able to predict, investigate and explain the


displacement–time, velocity–time and acceleration–time graphs for a
variety of motions, including trolleys on slopes. Consider the
uncertainties in each piece of equipment and which is most
appropriate for the measurements you are making. Evaluate your
experimental set-up to be able to suggest problems and potential
improvements.

Remember that displacement (s), velocity (u, v) and acceleration (a)


are vector quantities. A positive and negative direction must be chosen
and used consistently in your predictions and explanations.

8
The equations of motion
The equation which links acceleration, initial and final velocity, and
time is the first of the equations of motion.

These equations are used to describe motion in a straight line with


uniform acceleration. You must to be able to:

select the correct formula


identify the symbols and units used
carry out calculations to solve problems of real life motion
carry out experiments to verify the equations of motion.

You should develop an understanding of how the graphs of motion can


be used to derive the equations. This is an important part of
demonstrating that you understand the principles of describing
motion, and the link between describing it graphically and
mathematically.

a = acceleration in metres per second per second (m s –2 )


v u v = final velocity in metres per second (m s –1 )
a
t u = initial velocity in metres per second (m s –1 )
t = time in seconds (s)

v = u + at Equation of motion 1

9
You now should be sufficiently familiar with graphing motion to be
able to describe the motion represented by the graph below.

time (s)

This graph represents an object moving with a positive velocity


of 5 m s –1 , which is accelerating at a constant rate. After 300 s the
object is moving with velocity of 35 m s –1 . A constant acceleration
means the velocity is increasing at a constant rate.

The displacement of the object can be determined by calculating the


area under the graph. You may already be familiar with the idea of
using the area under a speed–time graph to determine the distance
travelled by an object.

10
Removing the numbers from the axis, we can work instead with the
symbols which represent the quantities, ie final velocity (v), initial
velocity (u) and time (t).

Notice that Equation 1 has been rearranged to give v – u = at and


substituted into the equation above.

Adding the two areas under the graph gives us:

s = displacement in metres (m)


2
s = ut + ½at u = initial velocity in metres per second (m s –1 )
t = time in seconds (s)
a = acceleration in metres per second per second (m s –2 )

s = ut + ½at 2 Equation of motion 2

11
The third equation of motion is derived from substituting Equation 1
into Equation 2.

Equation 1 v = u + at
square each side to give v 2 = (u + at) 2
v 2 = u 2 + 2uat + a 2 t 2
v 2 = u 2 + 2a(ut + ½at 2 )
substitute in Equation 2 v 2 = u 2 + 2as

v 2 = u 2 + 2as Equation of motion 3

12
Direction matters

Christine Arron was a 100-m sprint


athlete.

Immediately the starting pistol is


fired, Christine accelerates uniformly
from rest, reaching maximum velocity
at a distance of 21.8 m in 4.16 s.

Her maximum velocity is 10.49 m s –1 .

Calculate her acceleration over the


first 21.8 m of the race, showing full
working.

Using the normal sign convection in which right is positive and left is
negative, by calculation her acceleration is +2.52 m s –2 . In this case,
the positive value means increasing velocity with time in the positive
direction.

As she passes the finish line,


Christine begins to slow down.

She comes to rest in 8.20 s


from a velocity of 9.73 m s –1 .

Calculate her acceleration,


showing full working.

13
Using the normal sign convection in which right is positive and left is
negative, by calculation her acceleration is –1.19 m s –2 . In this case,
the negative value means decreasing velocity with time in the positive
direction.

Before continuing you should give some thought to what else a


positive or negative value of acceleration might mean.

Consider Christine running in the


opposite direction, where the sign
convention remains the same.

What will a positive value of


acceleration mean in this case? What
about a negative value?

Immediately the starting pistol is fired, Christine accelerates uniformly


from rest, in the opposite direction, reaching maximum velocity at a
distance of 21.8 m in 4.16 s.

Her maximum velocity is –10.49 m s –1 (why is it negative?).


Calculate her acceleration at a distance of 21.8 m of the race, showing
full working.

Her acceleration is –2.52 m s –2 . The negative value indicates that she is


gaining speed in the negative direction (or slowing down in the
positive direction).

As she passes the finish line, Christine begins to slow down. She comes
to rest in 8.20 s from a velocity of –9.73 m s –1 .
Calculate her acceleration, showing full working.

14
Her acceleration is a = 1.19 m s –2 . The positive value indicates that she
is losing speed in the negative direction.

When using equations in relation to motion, you must understand


what the values mean. Remember, equations are just one way of
describing motion – you should develop a picture in your head of the
actual motion being described by the mathematics.

Using the equations of motion: explaining good


technique using an example

Step 1: Write down the sign convention you are using for the situation.

Step 2: Write down what you know – think ssuvat

s sign convention (see step 1)


s displacement
u initial velocity
v final velocity
a acceleration
t time

Step 3: Write down any other information you have, e.g. acceleration
due to gravity.

Step 4: Write down your formulae and check against data sheet. Select
the most appropriate formula to use.

Step 5: Substitute values then rearrange the formula.

Step 6: Write answer clearly using magnitude with units and direction
(if appropriate).

15
Example
Usain Bolt is a Jamaican sprinter and a three-time Olympic gold
medallist (correct 2011).

Immediately the starting pistol is fired, Usain accelerates uniformly


from rest. He reaches 8.70 m s –1 in 1.75 s.

Calculate his displacement in this time.

Working

s  positive and  negative

s=?m u = 0 m s –1 v = 8.70 m s –1 a = ? m s –2 t = 1.75 s

(u is an easy one to miss – the phrase to look for is ‘starting from rest’)

Formula

v u at
1 2
s ut at
2
v2 u2 2as

s = (u + v)
2

Can this be done in one calculation? Is there one formula which links s,
v, u and t but does not require a?

On this occasion two formulae will be required, the first to determine


acceleration a and the second to calculate displacement s.

v = u + at
8.70 = 0 + a × 1.75
8.70 = 1.75a
a = 8.70
1.75
a = 4.97 m s –2

16
then

1 2
s ut at
2
1
s (0 t) 4.97 1.75 2
2
s 0 7.61
s = 7.61m

You should ensure that you are familiar with typical everyday
velocities and accelerations. This is key to understanding work in
physics on motion. For example, what is a realistic top speed for a
world-class sprinter?

What sort of accelerations do you experience in everyday life? Do you


experience motion only in the horizontal? An accelerometer (a device
which measures acceleration in three dimensions) can be used to
explore the accelerations that you experience during everyday
activities. Try it out – you might be surprised by the results!

17
Problems

Revision problems – Speed

1. The world downhill skiing speed trial takes place at Les Arcs
every year. Describe a method that could be used to find the
average speed of the skier over the 1 km run. Your description
should include:

any apparatus required


details of what measurements need to be taken
an explanation of how you would use the measurements to
carry out the calculations.

2. An athlete runs a 1500 m race in a time of 3 min 40 s. Calculate


his average speed for the race.

3. It takes light 8·0 minutes to travel from the Sun to the Earth.
How far away is the Sun from the Earth?
(speed of light = 3·0 × 10 8 m s 1 ).

4. The distance between London and New York is 4800 km. A plane
travels at an average speed of Mach 1·3 between London and
New York.
Calculate the time, to the nearest minute, for this journey.
(Mach 1 is the speed of sound. Take the speed of sound to be
340 m s 1 ).

5. The graph shows how the speed of a girl varies with time from
the instant she starts to run for a bus.

18
She starts from stand still at O and jumps on the bus at Q.
Find:

a) the steady speed at which she runs


b) the distance she runs
c) the increase in the speed of the bus while the girl is on it
d) how far the bus travels during QR
e) how far the girl travels during OR.

6. A ground-to-air guided missile starts from rest and accelerates


at 150 m s 2 for 5 s. What is the speed of the missile 5 s after
launching?

7. An Aston Martin has an acceleration of 6 m s 2 from rest. What


time does it take to reach a speed of 30 m s 1 ?

8. A car is travelling at a speed of 34 m s 1 . The driver applies the


brakes and the car slows down at a rate of 15 m s 2 . What is the
time taken for the speed of the car to reduce to 4 m s 1 ?

Revision problems – Acceleration

1. A skateboarder starting from rest goes down a uniform slope


and reaches a speed of 8 m s 1 in 4 s.

(a) What is the acceleration of the skateboarder?


(b) Calculate the time taken for the skateboarder to reach a
speed of 12 m s 1 .

2. In the Tour de France a cyclist is travelling at 16 m s 1 . When he


reaches a downhill stretch he accelerates to a speed of 20 m s 1
in 2·0 s.

(a) What is the acceleration of the cyclist down the hill?


(b) The cyclist maintains this constant acceleration. What is
his speed after a further 2·0 s?
(c) How long after he starts to accelerate does he reach a
speed of 28 m s 1 ?

19
3. A student sets up the apparatus shown to find the acceleration
of a trolley down a slope.

Length of card on trolley = 50 mm

Time on clock 1 = 0·10 s (time taken for card to interrupt top


light gate)
Time on clock 2 = 0·05 s (time taken for card to interrupt
bottom light gate)
Time on clock 3 = 2·50 s (time taken for trolley to travel
between top and bottom light gate)

Use these results to calculate the acceleration of the trolley.

20
Revision problems – Vectors

1. A car travels 50 km due north and then returns 30 km due


south. The whole journey takes 2 hours.

Calculate:

(a) the total distance travelled by the car


(b) the average speed of the car
(c) the resultant displacement of the car
(d) the average velocity of the car.

2. A girl delivers newspapers to three houses, X, Y and Z, as shown


in the diagram.

Z
N

40 m

30 m
X

She starts at X and walks directly from X to Y and then to Z.

(a) Calculate the total distance the girl walks.


(b) Calculate the girl’s final displacement from X.
(c) The girl walks at a steady speed of 1 m s 1 .
(i) Calculate the time she takes to get from X to Z.
(ii) Calculate her resultant velocity.

21
3. Find the resultant force in the following examples:

4. State what is meant by a vector quantity and scalar quantity.


Give two examples of each.

5. An orienteer runs 5 km due south then 4 km due west and then


2 km due north. The total time taken for this is 1 hour.

Calculate the average speed and average velocity of the


orienteer for this run.

6. A football is kicked up at an angle of 70° at 15 m s 1 .

Calculate:
(a) the horizontal component of the velocity
(b) the vertical component of the velocity.

22
Section 1: Equations of motion

Equations of motion

1. An object is travelling at a speed of 8·0 m s 1 . It then accelerates


uniformly at 4·0 m s 2 for 10 s. How far does the object travel in
this 10 s?

2. A car is travelling at a speed of 15·0 m s 1 . It accelerates


uniformly at 6·0 m s 2 and travels a distance of 200 m while
accelerating. Calculate the velocity of the car at the end of the
200 m.

3. A ball is thrown vertically upwards to a height of 40 m above its


starting point. Calculate the speed at which it was thrown.

4. A car is travelling at a speed of 30·0 m s 1 . It then slows down at


1·80 m s 2 until it comes to rest. It travels a distance of 250 m
while slowing down. What time does it take to travel the 250 m?

5. A stone is thrown with an initial speed 5·0 m s 1 vertically down


a well. The stone strikes the water 60 m below where it was
thrown.

Calculate the time taken for the stone to reach the surface of
the water. The effects of friction can be ignored.

6. A tennis ball launcher is 0·60 m long. A tennis ball leaves the


launcher at a speed of 30 m s 1 .

(a) Calculate the average acceleration of the tennis ball in


the launcher.
(b) Calculate the time the ball accelerates in the launcher.

23
7. In an experiment to find g a steel ball falls from rest through a
distance of 0·40m. The time taken to fall this distance is 0·29s.

What is the value of g calculated from the data of this


experiment?

8. A trolley accelerates uniformly down a slope. Two light gates


connected to a motion computer are spaced 0·50 m apart on the
slope. The speeds recorded as the trolley passes the light gates
are 0·20 m s 1 and 0·50 m s 1 .
(a) Calculate the acceleration of the trolley.
(b) What time does the trolley take to travel the 0·5 m
between the light gates?

9. A helicopter is rising vertically at a speed of 10·0 m s 1 when a


wheel falls off. The wheel hits the ground 8·00 s later.

Calculate the height of the helicopter above the ground when


the wheel came off. The effects of friction can be ignored.

10. A ball is thrown vertically upwards from the edge of a cliff as


shown in the diagram.

The effects of friction can be ignored.

(a) (i) What is the height of the ball above sea level 2·0 s
after being thrown?
(ii) What is the velocity of the ball 2·0 s after being
thrown?
(b) What is the total distance travelled by the ball from
launch to landing in the sea?

24
Motion – time graphs

1. The graph shows how the displacement of an object varies with


time.

displacement against time

North 8
7
displacement /m

6
5
4
3
2
1
0
0 1 2 3 4 5 6 7 8
time / s

(a) Calculate the velocity of the object between 0 and 1 s.


(b) What is the velocity of the object between 2 and 4 s from
the start?
(c) Draw the corresponding distance against time graph for
the movement of this object.
(d) Calculate the average speed of the object for the
8 seconds shown on the graph.
(e) Draw the corresponding velocity against time graph for
the movement of this object.

25
2. The graph shows how the displacement of an object varies wit h
time.

displacement against time

North 4.5
4
displacement /m

3.5
3
2.5
2
1.5
1
0.5
0
0 1 2 3 4 5 6
time / s

(a) Calculate the velocity of the object during the first


second from the start.
(b) Calculate the velocity of the object between 1 and 5 s
from the start.
(c) Draw the corresponding distance against time graph for
this object.
(d) Calculate the average speed of the object for the 5
seconds.
(e) Draw the corresponding velocity against time graph for
this object.
(f) What are the displacement and the velocity of the object
0·5 seconds after the start?
(g) What are the displacement and the velocity of the object
3 seconds after the start?

26
3. The graph shows the displacement against time graph for the
movement of an object.

displacement against time

North 2.5
2
1.5
displacement /m

1
0.5
0
-0.5 0 1 2 3 4 5
-1
-1.5
South -2
-2.5
time / s

(a) Calculate the velocity of the object between 0 and 2 s.


(b) Calculate the velocity of the object between 2 and 4 s
from the start.
(c) Draw the corresponding distance against time graph for
this object.
(d) Calculate the average speed of the object for the
4 seconds.
(e) Draw the corresponding velocity against time graph for
this object.
(f) What are the displacement and the velocity of the object
0·5 s after the start?
(g) What are the displacement and the velocity of the object
3 seconds after the start?

27
4. An object starts from a displacement of 0 m. The graph shows
how the velocity of the object varies with time from the start.

North velocity against time


–1
velocity m s
8
7
6
5
4
3
2
1
0
0 1 2 3 4 5 6 7 8
time / s time/s

(a) Calculate the acceleration of the object between


0 and 1 s.
(b) What is the acceleration of the object between 2 and 4 s
from the start?
(c) Calculate the displacement of the object 2 seconds after
the start.
(d) What is the displacement of the object 8 seconds after
the start?
(e) Sketch the corresponding displacement against time
graph for the movement of this object.

28
5. An object starts from a displacement of 0 m. The graph shows
how the velocity of the object varies with time from the start.

North
velocity against time
–1
velocity m s
2.5
2
North 1.5
1
0.5
0 time/s
-0.5 0 1 2 3 4 5
-1
-1.5
South -2
-2.5
time / s

(a) Calculate the acceleration of the object between


0 and 2 s.
(b) Calculate the acceleration of the object between
2 and 4 s from the start.
(c) Draw the corresponding acceleration against time graph
for this object.
(d) What are the displacement and the velocity of the object
3 seconds after the start?
(e) What are the displacement and the velocity of the object
4 seconds after the start?
(f) Sketch the corresponding displacement against time
graph for the movement of this object.

29
6. The velocity-time graph for an object is shown below.

velocity against time

15
velocity 10
/m s-1 5
0 time/s
-5 0 1 2 3 4 5 6 7

-10
time / s

A positive value indicates a velocity due north and a negative


value indicates a velocity due south. The displacement of the
object is 0 at the start of timing.

(a) Calculate the displacement of the object:


(i) 3 s after timing starts
(ii) 4 s after timing starts
(iii) 6 s after timing starts.

(b) Draw the corresponding acceleration–time graph.

7. The graph shows how the acceleration a of an object, starting


from rest, varies with time.

4
a
a/ms-2-2
ms 2

0 5 10
Time /
time/s
s

Draw a graph to show how the velocity of the object varies with
time for the 10 seconds of the motion.

30
8. The graph shows the velocity of a ball that is dropped and
bounces on a floor.
An upwards direction is taken as being positive.

+ C
Velocity O D
0
time
-
E
B

(a) In which direction is the ball travelling during section OB


of the graph?
(b) Describe the velocity of the ball as represented by
section CD of the graph.
(c) Describe the velocity of the ball as represented by
section DE of the graph.
(d) What happened to the ball at the time represented by
point B on the graph?
(e) What happened to the ball at the time represented by
point C on the graph?
(f) How does the speed of the ball immediately before
rebound from the floor compare with the speed
immediately after rebound?
(g) Sketch a graph of acceleration against time for the
movement of the ball.

31
9. A ball is thrown vertically upwards and returns to the thrower 3
seconds later. Which velocity-time graph represents the motion
of the ball?

10. A ball is dropped from a height and bounces up and down on a


horizontal surface. Which velocity-time graph represents the
motion of the ball from the moment it is released?

v/ms -1 v/ms -1 v/ms -1


-
v / m s- v / m s- v / ms
1 1 1
t/s
t/s C t/s
A B
-1
vv/ms
/ m s- -1
v / m s-
v/ms
1 1
0
t/s
D E t/s

32
11. Describe how you could measure the acceleration of a trolley
that starts from rest and moves down a slope. You are provided
with a metre stick and a stopwatch. Your description should
include:

(a) a diagram
(b) a list of the measurements taken
(c) how you would use these measurements to calculate the
acceleration of the trolley
(d) how you would estimate the uncertainties involved in the
experiment.

12. Describe a situation where a runner has a displacement


of 100 m due north, a velocity of 3 m s 1 due north and an
acceleration of 2 m s 2 due south. Your description should
include a diagram.

13. Is it possible for an object to be accelerating but have a


constant speed? You must justify your answer.

14. Is it possible for an object to move with a constant speed for 5


seconds and have a displacement of 0 m? You must justify your
answer.

15 Is it possible for an object to move with a constant velocity


for 5 s and have a displacement of 0 m? You must justify your
answer.

33
Solutions

Revision problems – Speed


1
2. 6·8 m s

3. 1·4 × 10 11 m

4. 181 minutes

5. (a) 5 ms 1
(b) 35 m
(c) 10 m s 1
(d) 100 m
(e) 135 m

1
6. 750 m s

7 5s

8. 2s

Revision problems – Acceleration


2
1. (a) 2 ms
(b) 6s

2. (a) 2·0 m s 2
(b) 24 m s 1
(c) 6·0 s

2
3. 0·20 m s

34
Revision problems – Vectors

1. (a) 80 km
(b) 40 km h 1
(c) 20 km north
(d) 10 km h 1 north

2. (a) 70 m
(b) 50 m bearing 037
(c) (i) 70 s
(ii) 0·71 m s 1 bearing 037

3. (a) 6·8 N bearing 077


(b) 11·3 N bearing 045
(c) 6·4 N bearing 129

5. Average speed = 11 km h 1
Average velocity = 5 km h 1 bearing 233

6. (a) 5·1 m s 1
(b) 14·1 m s 1

35
Section 1: Equations of motion

Equations of motion

1. 280 m

2. 51·2 m

1
3. 28 m s

4. 16·7 s

5. 3·0 s

2
6. (a) 750 m s
(b) 0·04 s

2 1
7. 9·5 m s or N kg

2
8. (a) 0·21 m s
(b) 1·4 s

9. 234 m

10. (a) (i) 21·4 m


1
(ii) 15·6 m s downwards
(b) 34·6 m

36
Motion–time graphs

1. (a) 2 m s 1 due north


(b) 0 ms 1
(d) 0·75 m s 1

2. (a) 4 m s 1 due north


(b) 1·0 m s 1 due south
(d) 1·6 m s 1
1
(f) displacement 2 m due north, velocity 4 m s due north
1
(g) displacement 2 m due north, velocity 1 m s due south

3. (a) 1 m s 1 due north


(b) 2 m s 1 due south
(d) 1.5 m s 1
1
(f) displacement 0·5 m due north, velocity 1 m s due north
(g) displacement 0, velocity 2 m s 1 due south

4. (a) 2 m s 2 due north


(b) 0 ms 2
(c) 4 m due north
(d) 32 m due north

5. (a) 1 m s 2 due north


(b) 2 m s 2 due south
1
(d) displacement 3 m due north, velocity 0 m s
1
(e) displacement 2 m due north, velocity 2 m s due south

6. (a) (i) 17·5 m due north


(ii) 22·5 m due north
(iii) 17·5 m due north

9. D. Note that in this question, downwards is taken to be the


positive direction for vectors.

10. A. Note that in this question, upwards is taken to be the positive


direction for vectors.

37
University of Nebraska - Lincoln
DigitalCommons@University of Nebraska - Lincoln

Robert Katz Publications Research Papers in Physics and Astronomy

1-1958

Physics, Chapter 16: Kinetic Theory of Gases


Henry Semat
City College of New York

Robert Katz
University of Nebraska-Lincoln, rkatz2@unl.edu

Follow this and additional works at: https://digitalcommons.unl.edu/physicskatz

Part of the Physics Commons

Semat, Henry and Katz, Robert, "Physics, Chapter 16: Kinetic Theory of Gases" (1958). Robert Katz
Publications. 166.
https://digitalcommons.unl.edu/physicskatz/166

This Article is brought to you for free and open access by the Research Papers in Physics and Astronomy at
DigitalCommons@University of Nebraska - Lincoln. It has been accepted for inclusion in Robert Katz Publications by
an authorized administrator of DigitalCommons@University of Nebraska - Lincoln.
16
Kinetic Theory of Gases

16-1 General Gas Law

The behavior of a gas under various condition8 of temperature and pressure


has already been studied in 80me detail. When the pressure of a constant
mass of gas is not too great, say less than about 2 atm, we find that a gas
obeys the following relationships:
at constant temperature PV = constant; (16-1)

at constant volume P = KT; (16-2)

at constant pressure V = K'T. (16-3)

These three equations are special cases of a single experimental equation


which gives the relationship between the pressure P, the volume V, and the

Fig. 16-1 Two steps in the derivation of the general gas law; an isothermal process
followed by a constant-pressure process.

absolute temperature T of a constant mass of gas. We may derive the


general form of the gas law from the above equations.
Let us consider a gas contained in a cylinder with a closely fitting
piston, as shown in Figure 16-1. The initial condition of the gas may be
299
300 KINETIC THEORY OF GASES § 16-1

described in terms of its initial pressure Pi, its initial volume Vi, and its
initial temperature T i . The gas is allowed to expand at constant tempera-
ture, say by keeping the cylinder immersed in a bath of melting ice, until
its new pressure is P f and its new volume is V 2 • Since the expansion was at
constant temperature, we find from Equation (16-1) that
PiV i = Pf V2 •

Now suppose that the gas is heated to a higher temperature T f , the volume
being allowed to expand to a new value Vj, but the pressure on the piston
being maintained at the same value P f throughout this process. Then,
from Equation (16-3) we may write

V 2 = V f = /('
Ti Tf '

Ti
or V 2 =Vf -·
Tf

Substituting for V 2 into the first of the above equations, we find

PiVi = PfVf .
(16.4)
Ti Tf

Equation (16-4) is one form of the general gas law. Since the initial state,
described by the subscript i, and the final state, described by the subscript
!, are entirely arbitrary, the only way in which the quantities on the right-
and left-hand sides of the equation can be equal is for each quantity to be
separately equal to the same constant. Thus we may rewrite the gas law as

~ (16-5)

~
where c is a constant whose value depends upon the mass of the enclosed
gas. Any convenient units may be used for the pressure and volume, but
the temperature T must always be the absolute temperature.
Illustrative Example. A given mass of air occupies a volume of 2,000 cm 3
at 27°C when its pressure corresponds to the pressure at the base of a column of
mercury 75 cm high. The air is compressed until its volume is 1,200 cm 3, and
its pressure corresponds to 225 cm of mercury. Determine the temperature of
the gas after it has been compressed.
From Equation (16-4) we have

PiV i = PfVf.
Ti Tf
§16-3 KINl''l'lC THl'ORY OF GASl'S 301

The pressure at the base of a column of mercury h em high is given by P = hpg,


where p is the density of the mercury. Substituting numerical values, we have
75 em X pg X 2,000 cm 3 225 em X pg X 1,200 cm 3 ,
300.2° abs Tf
from which 1j = 540.4° abs.

16-2 The Universal Gas Constant R

The constant c appearing in Equation (16-5) can be evaluated for any


given mass of a gas. Let us designate the value of this constant for a gram
molecular weight, or mole, of a gas by the symbol R. A gram molecular
weight of any substance is an amount of that substance whose mass, expressed
in grams, is numerically equal to the molecular weight of the substance. In the
limit of low pressures, the value of R is independent of the chemical nature
of the gas, so that R is known as the universal gas constant. In the event
that n moles of gas are present in a container, Equation (16-5) may be
rewritten as
PV = nRT. (16-6)

The numerical value of the gas constant R can be determined by noting


that 1 mole of any gas occupies a volume of 22.4 liters at a pressure of 76
cm of mercury at ooe; putting these values into Equation (16-6), we get
7 erg cal -2 liter atm
R = 8.31 X 10 mole °K = 1.99 mo1e oK = 8.21 X 10 mo e
1 oK

16-3 Kinetic Theory of Gases

From the preceding discussion we have seen that all gases exhibit similar
thermal and mechanical properties, regardless of their chemical composi-
tion, as long as their pressure is sufficiently small. This behavior is quite
unlike that of the same substances in liquid or solid form, where these
substances exhibit widely different thermal and elastic properties. We are
led to infer that the molecules of a gas are sufficiently far apart so that they
rarely interact with each other. The pressure of a gas then results from the
collisions of the molecules of the gas with the walls of the container. The
moving molecules of the gas completely fill every container in which the
gas is placed.
We may construct a theory of an ideal gas which is in good agreement
with the experimental results described in the preceding sections on the
basis of a few simple assumptions. We shall assume that a gas is composed
of molecules that are so small that, to a first approximation, they may be con-
sidered as point masses. We assume further that the molecules do not
302 KINETIC THEORY OF GASES §16-3

exert forces on each other except during collisions. We shall further assume
that the molecules of the gas are perfectly elastic, and that the container is
made of perfectly elastic, rigid walls. This implies that mechanical energy
is conserved in collisions between molecules. If this were not the case,
we would expect to observe that the pressure of a tank of gas would di-
minish with time, as the molecules lost mechanical energy in inelastic col-
lisions. For the sake of simplicity we shall assume that the gas is in a
cubical container of edge d and of volume V = d 3•

Fig. 16-2 Moleeules with equal


velocity components v 1 near face
BeDE of the cube.
.-J--:l----+---..",)D

The pressure exerted by the gas on the walls of the container is due to
the impact of the molecules on the walls, and, when in equilibrium, is
equal to the pressure throughout the gas. To calculate this pressure let
us assume that the impact of a molecule with a wall is an elastic impact;
that is, if a molecule is approaching the wall with a velocity v and mo-
mentum mv, then it will leave the wall with a velocity -v and a momentum
-mv. The change in momentum of the molecule produced by this impact
will thus be - 2mv. To determine the pressure on the walls of the container,
let us first calculate the force exerted by the molecules on one of the six
faces of the cube, say the face BCDE of Figure 16-2, and then divide by
its area.
Let us consider those molecules which at some instant are very close to
this face. Only those molecules whose velocities have components perpen-
dicular to this face, and directed toward it, will strike it and rebound. Sup-
pose we consider a small number of molecules which have the same value
VI for this velocity component. The number of these molecules which will
strike this face during a small time interval t:.t will be one half of the num-
ber contained in a small volume A t:.l, where A is equal to the area of the
face of the cube and t:.l = VI t:.t; the other half having a velocity component
of magnitude VI are moving away from the wall. If ni represents the num-
ber of molecules per unit volume which have a velocity component of
magnitude Vb then the number striking this face of the cube in time t:.t
will be
§16-3 KINETIC THEORY OF GASES 303

Since each such molecule wiII have its momentum changed by - 2mvi
as a result of this impact, the impulse imparted to the wall will be equal and
opposite to it, or +2mvl' The impulse F 1 t:.t on the wall produced by these
collisions in time t:.t will then be

from which
The pressure on the wall produced by the impact of these molecules is
F1 2
PI = A = nlmvl·

We can now consider another group of molecules, n2 per unit volume, which
have a slightly different velocity component V2 in this direction; they will
produce an additional pressure P2 given by
P2 = n2mv~.
In this way, we can break up the gas into different groups of molecules,
each group contributing a similar term to the pressure on this face of the
cube. The total pressure P due to all the different groups of molecules
will therefore be of the form
P = nl mvi + n2mv~ + n3mv~ + ....
This equation can be simplified by introducing a new term called the
average of the squares of the components of the velocities of all the molecules
moving perpendicular to face A and defined by the equation

nlvi + n2v~ + n3v~ +


n

in which n represents the total number of molecules per unit volume. Sub-
stituting this value of v~ in the equation for the pressure, we get
P = nmv~. (16-7)

There will be a similar expression for the pressure on each of the six
faces of the cube, except that the factor v~ will be replaced by the appro-
priate average of the squares of the components of the velocities of the
molecules for that particular face.
The velocity v of anyone molecule may be in any direction; it can be
resolved into three mutually perpendicular components Vx , V y , V z • The
magnitude of v in terms of the magnitudes of these components is given by
v = v; + v~ + v;.
2
304 KINETIC THEORY OF GASES §16-3

There will be a similar equation for the square of the velocity of each mole-
cule of the gas in terms of the squares of its three mutually perpendicular
components. If we add the squares of the component velocities in the x
direction and divide this sum by the total number of molecules, we will get
the average value of the square of this velocity component; it will be
represented by~. Similarly, ~ and ~ will represent the average squares
of the velocities in the y and z directions, respectively. By adding these
average squares of the three velocity components, we get
2v = v2
x
+2
vy + 2
v.,
where v2 is the average of the squares of the velocities of all the molecules.
Since the velocities of the molecules have all possible directions, the average
value of the squares of the velocity in anyone direction should be the same
as in any other direction, or

so that
If we take the x direction as perpendicular to the face A, we can write
v2 = 3v~,
so that Equation (16-7) becomes

(16-8)

Recalling that the kinetic energy of a moving molecule is equal to !mv 2 ,


Equation (16-8) may be written as
P = in(!mv 2 ).
Since n is the number of molecules per unit volume, we see that the pressure
is numerically equal to two thirds the kinetic energy of the molecules in a
unit volume of gas.
Let us suppose that No is the total number of molecules in a mole of
gas, called Avogadro's number, which is contained in a volume V. Then
the number of molecules per unit volume n is given by the expression
No
n=V'
Substituting for n into Equation (16-8), we find
PV = iN o X !mv 2 • (16-9)

Equation (16-9) is a theoretical result obtained from our hypotheses about


an ideal gas, relating the pressure and volume of 1 mole of an ideal gas.
§16-3 KINETIC THEORY OF GASES 305

If we compare this result to the experimental equation given in Equation


(16-6), which for 1 mole of gas becomes
PV = RT,
we find the theoretical and experimental results to be in agreement if
RT iNa !mv 2 ,
=

or if
1:2
-mv = -3 -R 7.,
2 2 Na

It is customary to define a new constant k, called Boltzmann's constant,


such that
R
k=-· (16-10)
Na
Since R is the gas constant per mole, and N a is the number of molecules in
a mole of gas, the constant k may be described as the gas constant per
molecule. In terms of k the preceding equation becomes
(16-11)

that is, the mean kinetic energy of translation of a molecule of gas is given
by ! the product of Boltzmann's constant by the absolute temperature.
This equation gives us some physical meaning of temperature for an ideal
gas. For such a gas the temperature is associated with the kinetic energy
of the random translational motion of the molecules of the gas. According
to Equation (16-11) the average energy of each molecule, and therefore the
total internal energy of an ideal gas, is associated with its temperature.
Thus the internal energy oj an ideal gas is a junction oj its temperature only,
and not of its pressure or volume.
In our derivation of the gas law in the form of Equation (16-9), we
used the word "molecules" to describe the particles with which we were
dealing. These molecules were described by the condition that they were
small, relatively far apart, and perfectly elastic. Thus this equation might
be used to describe the behavior of any aggregate of particles whose physical
dimensions were small compared to their average separation, provided that
these particles were elastic and rarely interacted with each other. The
neutrons in a nuclear reactor satisfy these conditions. If the neutrons are
in equilibrium with the material of the reactor at a temperature T, we
speak of them as thermal neutrons. The mean velocity v of these thermal
neutrons may be obtained from Equation (16-11). The particles of a
colloidal suspension may also be thought of as though they were molecules
of an ideal gas, and it is found that these also obey the gas laws.
Equation (16-9) incorporates another result called Avogadro's hy-
pothesis, first stated by Avogadro in 1811, that all gases occnpying equal
306 KINETIC THEORY OF GASES §16-4

volumes at the same temperature and pressure contain equal numbers of mole-
cules. The accepted value for the number of molecules in a mole of gas
No is
No = 6.023 X 1023 molecules/gm molecular wt.
As we have already seen, 1 mole of gas occupies a volume of 22.4 liters at
ooe and at a pressure of 76 em of mercury. If we perform the calculation
indicated in Equation (16-10) to find the numerical value of Boltzmann's
constant, we obtain
k = 1.38 X 10-16 erg;oK.

16-4 Work Done by a Gas

Whenever a gas expands against some external force, it does work on the
external agency; conversely, whenever a gas is compressed by the action
of some outside force, work is done on the gas. To calculate the work done

n
,I F=PA
(a) p ll
AII
,I
U

(b)
p

v
Fig. 16-3 (a) Expansion of a gas at constant pressure. (b) Graphical representation of
the work done !1Jr as an area on the PV diagram.

by a gas, consider a gas enclosed in a cylinder with a tight fitting piston.


The piston may be connected to a mechanical device on which it exerts
some force. The force F acting on the piston owing to the pressure P of
the gas is given by
F = PA,
in which A is the cross-sectional area of the piston, as shown in Figure 16-3.
Suppose that the piston is pushed out a small distance /lS, while the pressure
of the gas remains essentially constant. The work MY done by the gas in
§16-4 WORK DONE BY A GAS 307'

moving the piston is given by


tiJY = F tis = PA tis,
or (16-12)

Thus the work done by an expanding gas at constant pressure is equal to


the product of the pressure by the change in volume. No mechanical work
is done by a gas unless there is a change in the volume of the gas. The
first law of thermodynamics,
tiQ = tiU + tiJY, (15.6)

may be rewritten for processes involving gases as


tiQ = tiU + P tiV. (16-13)

Thus, in any process in which the volume of the gas remains constant,
called an isovolumic process, any heat delivered to the gas must appear as
internal energy and is therefore exhibited as a change in the temperature
of the gas.
Let us calculate the work done by an ideal gas which expands iso-
thermally, that is, at constant temperature, from an initial volume VI to a
final volume V 2 • From the gas law the relationship between the variables
of pressure, volume, and temperature for one mole of gas may be stated as
PV = RT.
The work done may be represented as an integral, from Equation (16-12), as

JY = fdJY = .£V2 P dV,


and, substituting for P its value from the gas law,
RT
P=-,
V
V2 dV
we find JY = RT
Ivv,
-
V
at constant temperature. Recalling that
dX
f --; = In x + C,
where C is a constant of integration, we find that
V2
JY = RTln- (16-14)
VI
for the work done by one mole of an ideal gas in an isothermal expansion
308 KINETIC THEORY OF GASES §16-5

at temperature T from an initial volume VI to a final volume V 2 • From


Equation (16-14) we see that when the gas expands, that is, when V 2 is
greater than Vb the work done is
P positive, as is consistent with the sign
convention developed in Section 15-6.
The work done is shown as the area
under the isotherm in Figure 16-4.

16-5 Molar Heat Capacity of a Gas

When a quantity of heat t:.Q is delivered


to a gas, it may change the internal
energy of the gas by an amount t:.U
V and may also result in the performance
of an amount of external work t:.Jr by
Fig. 16-4 Work done by a mole of a
the gas upon the outside world, in ac-
gas in an isothermal expansion at
temperature T from volume V 1 to cordance with the first law of ther-
volume V 2 • modynamics. If the volume of the
gas is kept constant, all the heat is
converted into internal energy. Since the internal energy U of a mole of
gas is a function of temperature only, we may define the molar heat capac-
ity at constant volume of a mole of gas C v as
C _ t:.Q _ t:.U
v - t:.T - t:.T (at constant volume), (16-15)

so that the change in internal energy t:.U may be expressed as


t:.U = C v t:.T. (16-16)

From Equation (16-13) the first law of thermodynamics as applied to an


ideal gas may be rewritten as
t:.Q = C v t:.T +p t:.V. (16-17)

Let us consider the change in temperature of a mole of gas when a


quantity of heat t:.Q is delivered to the gas while the pressure is held con-
stant but the volume is permitted to change. The thermal energy delivered
to the gas must now be used to do external work as well as to change the
internal energy of the gas. The rise in temperature of the gas will therefore
be less than in the case where the volume of the gas is kept constant. The
molar heat capacity at constant pressure Cp might be expressed in such
units as calories per mole per degree, and may be defined through the
equation
t:.Q
Cp = - (at constant pressure), (16.18)
t:.T
§16-5 MOLAR REA'!' CAPACITY OF A GAS 309

and from Equation (16-17) we find


Cv~T+P~V
Cp =
~T

From the gas law for a mole of gas we have


PV = RT,
R
or V = pl'.
At constant pressure both Rand P are constant, so that a change of volume
~ V is related to a change in temperature ~T through the equation

R
~V = p~T.

Substituting this result into the preceding equation for C p, we find that
Cp = C v + R. (16-19)

The molar heat capacity of a gas at constant pressure is always greater


than the molar heat capacity at constant volume by the gas constant R.
The internal energy of a monatomic gas, such as helium, is entirely in
the form of kinetic energy of translation of the random motions of the
atoms of the gas. From Section 16-3 this internal energy may be stated as
U = No X !mv 2 = !RT. (16-20)

Since R is a constant, we find that the change in internal energy ~U asso-


ciated with a change in temperature ~T is given by
~U =!R ~T.
Thus the molar heat capacity of a monatomic gas at constant volume may
be found by substituting the preceding result into Equation (16-15), to find
C v =!R (monatomic gas). (16-21a)

Substituting this result into Equation (16-19), we obtain


Cp = iR (monatomic gas). (16-21b)

It is customary to designate the ratio of the specific heat at constant pres-


sure to the specific heat at constant volume by the letter 'Y (gamma). Thus
Cp
'Y = - . (16.22)
Cv
Substituting from Equations (16-21) into (16-22), we find the value of 'Y
for a monatomic gas to be
'Y = 1 (monatomic gas).
310 KINETIC '.rHEORY OF GASES §16-5

In our development of the kinetic theory of an ideal gas we assumed


that a molecule could be considered as a point mass and showed that the
average kinetic energy of translation per molecule is ikT (Equation 16-11).
There are three independent directions of motion of translation, say the
x, y, and z directions. We say that the molecule has three degrees of freedom;
that is, three coordinates are necessary to specify the position of the mole-
cule at any instant, one for each degree of freedom. Since there is no reason
for preferring one direction rather than another, we postulate the principle
of equipartition of energy, that each degree of freedom should have the
same amount of energy. Referring to Equation 16-11, the amount of
energy to be associated with each degree of freedom per molecule is !kT.
The internal energy of a mole of a monatomic gas will then be iRT, as
given by Equation 16-20.
The idea of degrees of freedom can be extended to diatomic and poly-
tomic gases. It can be shown that the value of 'Y can be expressed as

7=--'
f+2
f
where f is the number of degrees of freedom per molecule. For a monatomic
3+2 5
gas we find that 'Y = - - = - = 1.67, in agreement with measured
3 3
values as shown in Table 16-1.
We may extend these ideas to a diatomic molecule which we may
imagine to be two point masses a fixed distance apart; the line joining the
two atoms is the axis of the molecule. If the diatomic molecule is con-
sidered as a rigid body, then it will have three degrees of freedom owing to
the translational motion of the entire molecule, plus a certain number of
degrees of freedom owing to the rotational motion of the molecule. A
glance at Table 16-1 shows that 'Y = 1.4 for diatomic gas, indicating that
f = 5; thus there must be two additional degrees of freedom of rotation.
These would correspond to rotations about two mutually perpendicular
axes in a plane at right angles to the line joining the two atoms.
Thus if at ordinary temperatures the molecules of a diatomic gas may
be thought of as rigid, and possessing no vibrational energy, the mean
energy of each molecule must be ~kT. The total internal energy of a mole
of such a gas is given by

U = No X ~kT = ~RT (diatomic gas). (16-23)

From this expression we find the molal' heat capacities of a diatomic gas
to be
C v = ~R, Cp = iR (diatomic gas), (16-24)
§16-6 ADIABATIC PROCESSES 311

and the ratio of the specific heats l' is


l' = i- (diatomic gas). (16-25)

These results are in rather remarkable agreement with experiment, as


shown in Table 16-1.
TABLE 16-1 THE MOLAR HEAT CAPACITY AT CONSTANT VOLUME,
AND THE RATIO OF SPECIFIC HEATS FOR SEVERAL GASES

Atoms per Cv* l'


Gas
Molecule Theory Experiment Theory Experiment
Argon 1 2.98 2.98 1.67 1.67
Helium 1 2.98 2.98 1.67 1.66
Oxygen 2 4.97 5.04 1.40 1.40
Nitrogen 2 4.97 4.93 1.40 1.40
Carbon monoxide 2 4.97 4.94 1.40 1.40

* The units of Cv are given as calories per mole per degree centigrade.

At high temperatures the classical theory of the specific heats of gases


is no longer adequate to describe the experimental determinations. Addi-
tional rotational and vibrational modes are excited at high temperatures
in a manner which is best described by the quantum theory of specific
heats; this is beyond the scope of this book.

16-6 Adiabatic Processes


An adiabatic process is one in which no heat enters or leaves the system.
In the form of an equation
flQ = O.
If we apply the first law of thermodynamics, m the form of Equation
(16-17), to adiabatic processes, we find
flQ = 0 = C v flT + P flV.
flV
Thus flT = -p--.
Cv
The general gas law describes the behavior of a gas under all circumstances.
Thus for a mole of gas we have
PV = RT.
If P, V, and T are permitted small variations, we find, on taking differ-
entials, that
P fl V + V flP = R flT.
312 KINETIC THEORY OF GASES §16-6

If we replace the gas constant R by


R = Cp - Cy

from Equation (16-19), and substitute the value of t:,.T into the above
equation, we find
Pt:,.V
P t:,.V + V t:,.P = -(C p - C y ) - - = -("I - 1) P t:,.V,
Cy
for, from Equation (16-22), "I = CpjC y . On transposing, and dividing
the above equation by the product PV, we find
t:,.P t:,.V
p+'Y y = O.

In the limit of small increments we may replace t:,. by d, and integrate to


find
In P + "I In V = constant,
or PV'Y = constant. (16-26)

The value of the constant is determined by the quantity of gas present, so


that Equation (16-26) may be used to describe the relationship between
the pressure and the volume of any quantity of gas undergoing an adiabatic
change. If a gas originally at pressure PI, volume VI, and temperature T I
is compressed adiabatically, as in an insulated cylinder, to a new pressure
P z , the new volume of the gas V z may be found from Equation (16-26) by
writing
(16-27)

The final temperature of the gas T z may then be obtained from the gas law
PI VI PzV z
--=--,
TI Tz
in which all quantities except T z are now known.
Illustrative Example. A mass of gas occupies a volume of 8 liters at a pres-
0
sure of 1 atm and a temperature of 300 abs. It is compressed adiabatically to
a volume of 2 liters. Determine (a) the final pressure and (b) the final temperature,
assuming it to be an ideal gas whose value of "I = 1.5.
(a) The final pressure of the gas can be determined with the aid of Equation
(10-27) thus,

so that

from which
§16-7 'l'HE MAXWELL DISTRIBUTION FUNCTION 313

(b) The final temperature can be found with the aid of the general gas law
thus,
T2 = Tl P 2V2 ,
PlV l
0 b 8 atm X 21
so that T2 = 300 a s ,
1 atm X 81
or T2 = 600 0 abs.

16-7 The Maxwell Distribution Function

In Section 16-3 we showed how the properties of a gas could be accounted


for on the basis of a very simple set of hypotheses about the nature of a
gas. We assumed that the gas was made up of many molecules in rapid
motion, and that the molecules were sufficiently far apart so that the forces

3
Speed
Fig. 16-5 The Maxwellian distribution of molecular speeds. Relative numbers of
molecules having speeds in a unit speed interval at various speeds are shown as ordinate,
while speeds in units of the most probable speed are shown as abscissa.

that one molecule exerted on another were of minor importance and could
be neglected. We assumed that the molecules were perfectly elastic so
that there was no loss in mechanical energy in collisions between molecules
of the gas and the walls of the container. On the basis of such arguments,
we could account for the gas law, and we were able to show that the tem-
perature of a gas was directly related to the average value of the kinetic
energy of its molecules.
When a gas is in equilibrium at an absolute temperature T, the dis-
tribution of velocities of the molecules of the gas is given by Figure 16-5,
called the Maxwellian distribution, according to the theory first developed
by Maxwell (1831-1879). This theory has now been well verified by experi-
ment as actually describing the behavior of gas molecules. In fact, the
Maxwellian distribution may be taken as the meaning of the temperature
314 KINETIC THEORY OF GASES §16-7

of a gas, for, if a collection of gas molecules has a velocity distribution


which differs from Figure 16-5, then we may say that the gas has not yet
reached thermal equilibrium and therefore does not have a well-defined
temperature.
The average of the velocity components of the molecules of a gas in
a particular direction must be zero. If this were not so, the gas and its
container would be in translational motion. However, the average value
of the squares of the molecular velocities is not zero, and is given by Equa-
tion (16-11). From Figure 16-5 we see that molecules whose speeds are
more than three times the most probable speed are extremely rare. Never-
theless, there are some molecules in the gas which have very large speeds,
for the distribution curve approaches the horizontal axis asymptotically.
We must also note that, at a given temperature, the molecules of a gas of
low molecular weight are in more rapid motion than the molecules of a gas
of high molecular weight. This has the interesting consequence that
hydrogen and helium are steadily diffusing out of the earth's atmosphere,
for, at the temperature of the outer air, some of these lighter molecules are
moving sufficiently rapidly to attain the escape velocity of 11 km/sec
necessary for a projectile to escape the gravitational pull of the earth.
If we call the energy required to disrupt a chemical molecule its binding
energy, we see that, as the temperature of a gas is raised, a greater propor-
tion of gas molecules may have kinetic energies greater than the binding
energy, so that a molecule may be decomposed as a result of energy transfer
during a collision. Thus molecules which are stable at ordinary tempera-
tures must have binding energies which are large compared to the mean
kinetic energy of a molecule at room temperature, as given by Equation
(16-11).
It is interesting that modern theories of the structure of atoms and
molecules have provided a justification of a basic assumption of the kinetic
theory of gases. According to the quantum theory, molecules exist only in
certain quantum states, each having a fixed amount of energy. These are
sometimes called energy levels. The molecule normally exists in its state
of lowest energy, called its ground state, and can only absorb energy in a
collision with another molecule in exactly the right amount to raise it to a
state of higher energy, called an excited state, or to disrupt it completely.
In general, it is very unlikely that the colliding molecules will have just the
right amount of energy for excitation, so that the collisions between the
molecules of a gas result in no absorption of energy by the molecules. The
kinetic energy is conserved in the collision rather than being transferred
to internal excitation energy of one molecule. The collisions are therefore
perfectly elastic.
PROBLEMS 315

Problems
16-1. A closed vessel contains dry air at 25°C and 76 cm of mercury pressure.
Its temperature is raised to 100°C. Determine the pressure of the air, neglecting
the change in volume of the container.
16-2. A mass of oxygen occupies a volume of 1 liter at a pressure of 76 cm
of mercury when its temperature is 40°C. The gas is allowed to expand until its
volume is 1.5 liters and its pressure is 80 cm of mercury. (a) Determine its final
temperature. (b) Determine the number of moles of oxygen in the system.
16-3. Derive the general gas law from Equations (16-1) to (16-3) by con-
sidering that the gas is taken from TiP i Vi by a constant-volume process to
T 2Pf Vi, and thence by a constant-pressure process to TfPf Vf'
16-4. A certain gas has a density of 0.001 gm/cm 3 when its temperature is
50°C and its pressure is 4 atm. What pressure will be needed to change the
density of the gas to 0.002 gm/cm 3 when its temperature is 100°C?
16-5. An automobile tire has a volume of 1,000 in. 3 and contains air at a
gauge pressure of 24 Ib/in. 2 when the temperature is ODC. ,"Vhat will be the
gauge pressure of the air in the tires when its temperature rises to 27 DC and its
volume increases to 1,020 in. 3 ?
16-6. Determine the pressure of 4.032 gm of hydrogen which occupies a
volume of 16.8 liters at a temperature of O°C. The molecular weight of-hydrogen
is 2.016.
16-7. Determine the average value of the kinetic energy of the molecules
of a gas (a) at ODC and (b) at 100°C.
16-8. (a) What is the mass of a hydrogen molecule? (b) Determine the
average velocity of a molecule of hydrogen at 27 DC.
16-9. The molecules of a certain gas have a mass of 5 X 10-24 gm. What
is the number of molecules per cubic centimeter of this gas when its pressure is
10 6 dynes/cm 2 and its temperature is 27 DC?
16-10. Calculate the work done in compressing one mole of oxygen from a
volume of 22.4 liters at ODC and 1 atm pressure to 16.8 liters at the same tem-
perature.
16-11. A cylinder contains a mole of hydrogen at ODC and 76 cm of mercury
pressure. Calculate the amount of heat required to raise the temperature of this
hydrogen to 50 DC (a) keeping the pressure constant, and (b) keeping the volume
constant. (c) What is the volume of the hydrogen when at ODC?
16-12. A cylinder contains 32 gm of oxygen at ODC and 76 cm of mercury
pressure. Calculate the amount of heat required to raise the temperature of this
mass of oxygen to 80DC (a) keeping the pressure constant and (b) keeping the
volume constant. (c) How much mechanical work is done by the oxygen in
each case?
16-13. A mass of a monatomic gas occupies a volume of 400 cm 3 at a tem-
perature of l7°C and a pressure of 76 cm of mercury. The gas is compressed
adiabatically until its pressure is 90 cm of mercury. Determine (a) the final
volume of the gas and (b) the final temperature of the gas.
16-14. A mass of a diatomic gas occupies a volume of 6 liters at a temperature
316 KINETIC THEORY OF GASES

of 27°C and 75 cm of mercury pressure. The gas expands adiabatically until its
volume is 8 liters. What is the final temperature of the gas?
16-15. A mole of gas at atmospheric pressure and O°C is compressed iso-
thermally until its pressure is 2 atm. How much mechanical work is done on
the gas during this operation?
16-16. Ten grams of oxygen are heated at constant atmospheric pressure
from 27°C to 127°C. (a) How much heat is delivered to the oxygen? (b) What
fraction of the heat is used to raise the internal energy of the oxygen?
16-17. An air bubble of volume 20 cm 3 is at the bottom of a lake 40 m deep
where the temperature is 4°C. The bubble rises to the surface where the temper-
ature is 20°C. Assuming that the temperature of the bubble is the same as
that of the surrounding water, what is its volume just as it reaches the surface?
16-18. An ideal gas for which l' = 1.5 is enclosed in a cylinder of volume
1 m 3 under a pressure of 3 atm. The gas is expanded adiabatically to a pressure
of 1 atm. Find (a) the final volume and (b) the final temperature of the gas if
its initial temperature was 20°C.
16-19. A mass of 1.3 kg of oxygen of molecular weight 32 is enclosed in a
cylinder of volume 1 m 3 at a pressure of 10 5 nt/m 2 and a temperature of 20°C.
From these data find the universal gas constant R assuming oxygen to be an
ideal gas.
16-20. A gas of mass m and molecular weight M undergoes an isothermal
expansion from an initial pressure PI and volume V I to a final pressure P 2 and
volume V 2 while at temperature T. Find (a) the work done by the gas in this
expansion, (b) the heat flow to the gas, and (c) the change in internal energy of
the gas in terms of these symbols.
16-21. A piece of putty is placed in a vise with insulating jaws. A constant
force of 100 nt is applied through a distance of 2 cm. The putty is found not to
have its volume changed in this process. What is the change in the internal
energy of the putty?
16-22. Prove that TV'Y- I = constant for an adiabatic process.
16-23. Show that the work done by a gas in an adiabatic expansion from
initial conditions Pi, Vi to final conditions P 1> V f is given by

}f' = Pi Vi (V}-'Y _ V~-'Y).


1 - l'
Chapter 3

Static Equilibrium

3.1 The Important Stuff


In this chapter we study a special case of the dynamics of rigid objects covered in the last
two chapters. It is the (very important!) special case where the center of mass of the object
has no motion and the object is not rotating.

3.1.1 Conditions for Equilibrium of a Rigid Object


For a rigid object which is not moving at all we have the following conditions:
• The (vector) sum of the external forces on the rigid object must equal zero:
X
F=0 (3.1)

When this condition is satisfied we say that the object is in translational equilibrium.
(It really only tells us that aCM is zero, but of course that includes the case where the object
is motionless.)
• The sum of the external torques on the rigid object must equal zero.
X
τ =0 (3.2)

When this condition is satisfied we say that the object is in rotational equilibrium. (It
really only tells us that α about the given axis is zero, but —again— that includes the case
where the object is motionless.)
When both 3.1 and 3.2 are satisfied we say that the object is in static equilibrium.
Nearly all of the problems we will solve in this chapter are two–dimensional problems (in
the xy plane), and for these, Eqs. 3.1 and 3.2 reduce to
X X X
Fx = 0 Fy = 0 τz = 0 (3.3)

55
56 CHAPTER 3. STATIC EQUILIBRIUM

3.1.2 Two Important Facts for Working Statics Problems


i) The force of gravity acts on all massive objects in our statics problems; its acts on all the
individual mass points of the object. One can show that for the purposes of computing the
forces and torques on rigid objects in statics problems we can treat the mass of the entire
object as being concentrated at its center of mass; that is, for an object of mass M we can
treat gravity as exerting a force Mg downward at the center of mass.
(This result depends on the fact that the acceleration of gravity, g is usually constant
over the volume of the object. Otherwise it is not true.)

ii) While there is only one way to write the conditions for the forces on a rigid object
summing to zero, we have a choice in the way we write the equation for the total torque.
Eq. 3.3 does not specify the choice of the axis for calculating the torque. In general it matters
a great deal which axis we pick! But when the sum of torques about any one axis is zero
and the sum of forces is zero (translational equilibrium) then the sum of torques about any
axis will give zero; so for statics problems we are free to pick the most convenient axis for
P
computing τ . Often this will be the point on the object where several unknown forces are
acting, so that the resulting set of equations will be simpler to solve.

3.1.3 Examples of Rigid Objects in Static Equilibrium


Strategy for solving problems in static equilibrium:
• Determine all the forces that are acting on the rigid body. They will come from the other
objects with which the body is in contact (supports, walls, floors, weights resting on them)
as well as gravity,
• Draw a diagram and put in all the information you have about these forces: The points
on the body at which they act, their magnitudes (if known), their directions (if known).
• Write down the equations for static equilibrium. For the torque equation you will have a
choice of where to put the axis; in making your choice think of which point would make the
resulting equations the simplest.
• Solve the equations! (That’s not physics. . . that’s math.) If the problem is well–posed you
will not have too many or too few equations to find all the unknowns.

3.2 Worked Examples


3.2.1 Examples of Rigid Objects in Static Equilibrium

1. The system in Fig. 3.1 is in equilibrium with the string in the center exactly
horizontal. Find (a) tension T1, (b) tension T2, (c) tension T3 and (d) angle θ.
[HRW5 13-23]
3.2. WORKED EXAMPLES 57

35o
T1 q T3

T2
40 N 50 N

Figure 3.1: System of masses and strings for Example 1.

35o
T3
T1 q
T2 T2

40 N 50 N

(a) (b)

Figure 3.2: (a) Forces at the left junction of the strings. (b) Forces acting at the right junction of the
strings.

Whoa! Four unknowns (T1 , T2 , T3 and θ) to solve for! How will we ever figure this out?
We consider the points where the strings meet; the left junction is shown in Fig. 3.2 (a).
Since a string under tension pulls inward along its length with a force given by the string
tension, the forces acting at this point are as shown.
Since this junction in the strings is in static equilibrium, the (vector) sum of the forces
acting on it must give zero. Thus the sum of the x components of the forces is zero:
−T1 sin 35◦ + T2 = 0 (3.4)
and the sum of the y components of the forces is zero:
+T1 cos 35◦ − 40 N = 0 (3.5)
Now we look at the right junction of the strings; the forces acting here are shown in
Fig. 3.2 (b). Again, the sum of the x components of the forces is zero:
−T2 + T3 sin θ = 0 (3.6)
and the sum of the y components of the forces is zero:
+T3 cos θ − 50 N = 0 (3.7)
58 CHAPTER 3. STATIC EQUILIBRIUM

And at this point we are done with the physics because we have four equations for four
unknowns. We will do algebra to solve for them.
In this problem the algebra really isn’t so bad. From Eq. 3.5 we get
(40 N)
T1 = = 48.8 N
(cos 35◦ )
and then Eq. 3.4 gives us T2 :

T2 = T1 sin 35◦ = (48.8 N) sin 35◦ = 28.0 N .

We now rewrite Eq. 3.6 as:

T3 sin θ = T2 = 28.0 N (3.8)

and Eq. 3.7 as:


T3 cos θ = 50.0 N (3.9)
Now if we divide the left and right sides of 3.8 by the left and right sides of 3.9 we get:
(28.0 N)
tan θ = = 0.560
(50.0 N)
and then
θ = tan−1 (0.560) = 29.3◦
Finally, we get T3 from Eq. 3.9:
(50.0 N)
T3 = = 57.3 N
(cos 29.3◦ )
Summarizing, we have found:

T1 = 48.8 N T2 = 28.0 N T3 = 57.3 N θ = 29.3◦

This answers all the parts of the problem.

2. The system in Fig. 3.3 is in equilibrium. A mass of 225 kg hangs from the end
of the uniform strut whose mass is 45.0 kg. Find (a) the tension T in the cable
and the (b) horizontal and (c) vertical force components exerted on the strut by
the hinge. [HRW5 13-33]

(a) The rigid body here is the strut. What are the forces acting on it?
We know the mass M of the strut; the force of gravity exerts a force Mg downward at
its center of mass (which is in the middle of the strut since it is uniform). If the hanging
mass is m = 225 kg then the string which supports it exerts a downward force of magnitude
mg at the top end of the strut. The cable attached to the top of the strut exerts a force
of magnitude T . What is its direction? Some geometry (shown in Fig. 3.4 ) shows that its
3.2. WORKED EXAMPLES 59

Strut

T
225
kg

30o 45o
Hinge
Figure 3.3: Geometry of the statics problem of Example 2.

T
o
15
mg
Fh

Mg
30o 135
o
45o
Hinge
Figure 3.4: Forces acting on the strut in Example 2.
60 CHAPTER 3. STATIC EQUILIBRIUM

direction makes an angle of 15◦ with the strut. Finally the hinge exerts a force on the strut.
(Can’t forget that. . . the hinge is in contact with the metal bar which is the “strut” as so
exerts a force on it.) The magnitude of this force is just labelled Fh in the diagram, but we
don’t know its direction!
Now, one way to solve the problem would be to let the direction of the hinge force be
some angle θ as measured from some line of reference. In fact it will probably be easiest to
let the x and y components of this force be the unknowns... I will call them Fh,x and Fh,y . In
fact, parts (b) and (c) of the problem ask us for these components directly. We can always
get the direction and magnitude later!
Now let’s write down some equations. First, the sum of the Fx ’s must give zero. Note
(from basic geometry) that the force of the cable is directed at 30◦ below the horizontal.
And force of the hinge has an x component! Then from Fig. 3.4 we immediately read off:
Fh,x − T cos 30◦ = 0 (3.10)
Good enough. Now onto the Fy equation. The sum of the y components of the forces
gives zero, and we write:
Fh,y − T sin 30◦ − Mg − mg = 0 (3.11)
Now we use the condition for zero net torque. The question is: Where do we want to
put the axis? For this problem, the answer is obvious. We want to put it at the hinge
itself because then when we calculate the torques, the hinge force (with its two unknown
components) will give no torque. The equation will still be useful. . . and it will be much
simpler. (Keep in mind that even though a physical strut really does turn around a physical
hinge we still have the choice of putting the axis for torque anywhere.)
We are not told the length of the strut, so let its length be L. We note the angles that
the force vectors make with the line joining the axis to the points of application, and then
we write the sum of the torques as:
L
−Mg sin 45◦ − mgL sin 45◦ + T L sin 15◦ = 0
2
but we note that we can cancel the L out of this equation, leaving
Mg
− sin 45◦ − mg sin 45◦ + T sin 15◦ = 0 (3.12)
2
Are we done with the physics yet? In the Eqs. 3.10, 3.11 and 3.12 there are three
unknowns: T , Fh,x and Fh,y . We are done with the physics. Only algebra remains.
And the algebra isn’t so bad. The only unknown in Eq. 3.12 is T and we get:
Mg
T sin 15◦ = sin 45◦ + mg sin 45◦
2
= 12 (45 kg)(9.80 sm2 ) sin 45◦ + (225 kg)(9.80 sm2 ) sin 45◦
= 1715 N
so that
(1715 N)
T = = 6.63 × 103 N
sin 15◦
3.2. WORKED EXAMPLES 61

L
m
M=2m CM

60o xL L/2

Figure 3.5: Geometry of the statics problem of Example 3. Student is standing on a ladder which leans
against a wall.

(b) With our result from part(a) in hand, Fh,x and Fh,y will be easy to find. From 3.10 we
get:
Fh,x = T cos 30◦ = (6630 N) cos 30◦ = 5.74 × 103 N
and From 3.11 we get:

Fh,y = T sin 30◦ + Mg + mg


= (6630 N) sin 30◦ + (45 kg)(9.80 sm2 ) + (225 kg)(9.80 sm2 )
= 5.96 × 103 N

The horizontal and vertical components of the force of the hinge on the strut are

Fh,x = 5740 N Fh,y = 5960 N .

3. A ladder having a uniform density and a mass m rests against a frictionless


vertical wall at an angle of 60◦ . The lower end rests on a flat surface where the
coefficient of static friction is µs = 0.40. A student with a mass M = 2m attempts
to climb the ladder. What fraction of the length L of the ladder will the student
have reached when the ladder begins to slip? [Ser4 12-13]

We make a basic diagram of the geometry of the problem in Fig. 3.5. The ladder has
length L; we show the center of mass of the ladder at a distance L2 up from its bottom end.
If the student had climbed a fraction x of the ladder, then he/she is at a distance xL from
its lower end, as shown.
Keeping the geometry in mind, we next think about all the separate forces that are acting
on the ladder as it leans against the wall and supports the student.
The force of gravity mg (downward) is effectively exerted at the center of the ladder. Since
the ladder is exerting an upward force Mg on the student, the student must be exerting a
62 CHAPTER 3. STATIC EQUILIBRIUM

Nw

Nf CM

mg
Mg fs

Figure 3.6: Forces acting on the ladder in Example 3. Many TTU students do look like this guy.

downward force of magnitude Mg = 2mg on the ladder at a point a distance xL from the
lower end. The wall is frictionless so it can only exert a normal force on the top end of the
ladder; we will denote the magnitude of this force by Nw . The floor will exert a normal force
Nf upward on the bottom end of the ladder but also a horizontal force of static friction.
Which way does this friction force point? In our diagram, the wall’s normal force points to
the left so the friction force must point to the right so that the forces can add up to zero.
All these forces and their directions are diagrammed in Fig. 3.6. Now we apply the
conditions for static equilibrium given in Eq. 3.3.
First off, the horizontal forces must sum to zero. That gives us:
fs − Nw = 0 (3.13)
Next, the vertical forces must sum to zero. This gives:
Nf − Mg − mg = 0
and using M = 2m we get:
Nf − 3mg = 0 (3.14)
or: Nf = 3mg, giving us an expression for the normal force of the floor.
The next condition for equilibrium is that the sum of torques taken about any axis must
give zero. Since we have two forces acting at the lower end of the ladder, it might be best to
put the axis there because then those forces will give no torque, and we will have a simpler
equation to deal with. We note that the gravity forces from the student and the ladder’s
CM make an angle of 30◦ with the line joining the axis to the application points; they give
a clockwise (negative) torque. The normal force from the wall makes an angle of 60◦ with
the line from the axis, and it gives a positive torque. Our equation is:
−(xL)(2mg) sin 30◦ − (L/2)(mg) sin 30◦ + (L)(Nw ) sin 60◦ = 0
This equation can be simplified: We can cancel an L and use the values of sin 30◦ and sin 60◦
to get: √
mg 3
−xmg − + Nw = 0 (3.15)
4 2
3.2. WORKED EXAMPLES 63

We now have three equations, with four unknowns Nf , Nw , fs and x. (We can consider
m as given. In fact, its value won’t matter.) We need another equation!
We have not yet used the condition that when the student is standing xL up from the
ladder’s bottom it is just about to slip. Why should the ladder slip at all? It is because the
force of static friction fs is limited in size; we know that it can only be as large as µs Nf , since
Nf is the normal force between the floor and the ladder’s lower end. When the student has
walked up far enough that the ladder is on the verge of slipping then we have the equality

fs = µs Nf (3.16)

That’s all the equations! Now let’s start solving them. (The physics is done. The math
remains.)
Using Eq. 3.16 in Eq. 3.13 gives

µs Nf − Nw = 0

but from 3.14 we had Nf = 3mg, so we get

3µs mg = Nw

Put this result in Eq. 3.15 and we have:



mg 3
−xmg − + (3µs mg) = 0
4 2
We can cancel out the factor mg since it appears in each term. (So we never needed to know
m): √
1 3 3µs
−x − + =0
4 2
And at last, using the given value µs = 0.400:

3 3(0.400) 1
x= − = 0.798
2 4
The student can climb a fraction of 0.789 (that is, nearly 80%) of the length of the ladder
before it starts to slip.

4. For the stepladder shown in Fig. 3.7, sides AC and CE are each 8.0 ft long
and hinged at C. Bar BD is a tie–rod 2.5 ft long, halfway up. A man weighing
192 lb climbs 6.0 ft along the ladder. Assuming that the floor is frictionless and
neglecting the weight of the ladder, find (a) the tension in the tie–rod and the
forces exerted on the ladder by the floor at (b) A and (c) E. Hint: It will help
to isolate parts of the ladder in applying the equilibrium conditions. [HRW5 13-41]

Before anything else, we need to do a little geometry to find the angles which the ladder
sides make with with floor, because we are given enough information for this in the problem.
64 CHAPTER 3. STATIC EQUILIBRIUM

B D

A E

Figure 3.7: Man stands on a ladder in Example 4.

T
f Nf
ft

192 lb
4.0

q q
1.25 ft

(a) (b)

Figure 3.8: (a) Some trigonometry to find the angle of slope of the ladder. (b) The forces which act on
the left side of the ladder and their application points.
3.2. WORKED EXAMPLES 65

We note that since the sides are of equal length then triangle ACE is an isosceles triangle.
The part of the ladder above the tie–rod has the lengths shown in Fig. 3.8(a). From this we
can see that the angle which either side of the ladder makes with the ground is the same as
the angle θ shown in the figure, and θ is given by
1.25
cos θ = = 0.3125 =⇒ θ = 71.79◦
4.00
and the angle φ (which we’ll also need) is the complement of θ and is given by

φ = 90◦ − θ = 18.21◦

Now that the geometry is settled, we follow the hint and treat the sides of the ladder as
separate objects for which we will apply the conditions of static equilibrium. First, consider
the left side of the ladder (the one on which the man is standing). What forces are acting
on it?
The floor is frictionless (!) and so it can only exert a normal (vertical) force on the lower
end. We will let this force be Nf . Moving up, the tie-bar is under a tension T , so it exerts
a force of magnitude T which (we would guess) is pulling inward along its length, and so
points to the right. This force is applied at the midpoint of the side. The 192 lb man is
supported by the ladder, so he must be exerting a downward force of 192 lb at the point that
is 6 ft up from the lower end.
But we’re not done. At point C, the right side of the ladder is exerting a force on the left
side (the one we are now considering). This is true because the two parts of the ladder are
in contact at this point. Right now it is not at all clear which direction this force will point,
so we will just say it is some force F with components Fx and Fy which we will determine
in the course of solving the equations. This force is indicated in the figure. Now we have all
the forces acting on the left side of the ladder; they are diagrammed in Fig. 3.8 (b).
Thinking ahead to when we do our analysis for the right side of the ladder. . . . we note
that from Newton’s 3rd law the sides of the ladder will exert “equal and opposite” forces on
each other at point C. So the force of the left side on the right side will be −F. But that
comes later.
The sum of the x forces on this part of the ladder gives zero. That gives us:

Fx + T = 0 (3.17)

The sum of the y forces on this part of the ladder gives zero:

Fy − 192 lb + Nf = 0 (3.18)

Next, the sum of the torques on the left ladder section must give zero. We will choose
the location for our axis to be at the top (that is, point C); the reason for this choice is that
the unknown force F (which acts at C) will then give no torque. Using the given dimensions
of the ladder and the application points of the forces (as well as the angles θ and φ that we
figured out), we find:

−(8 ft)Nf sin φ + (2 ft)(192 lb) sin φ + (4 ft)T sin θ = 0 (3.19)


66 CHAPTER 3. STATIC EQUILIBRIUM

-F

T N’f

Figure 3.9: The forces which act on the right side of the ladder.

And now we analyze right side of the ladder. Starting from the bottom, we have the
force of the floor. The floor is frictionless, so again the force points straight up, but it will
not have the same magnitude as on the left side; we denote its magnitude here by Nf0 . Then,
midway up the length of the ladder the tension of the tie–rod pulls with a force of magnitude
T to the left. Finally at point C, the left side of the ladder is exerting a force. But it must
be opposite to the force we already assigned as coming from the right side. So the left side
exerts a force −F on the right side of the ladder. These forces are diagrammed in Fig. 3.9.
And now we get the equations. The sum of the x forces on this part of the ladder gives
zero:
−Fx − T = 0
But note that this equation is basically the same as Eq. 3.17. So we can ignore it!
The sum of the y forces on this part of the ladder gives zero:
−Fy + Nf0 = 0 (3.20)
Finally, the sum of the torques on the left ladder section must give zero. Again we choose
the top (point C) as the location for our axis (for the same reasons as before). Using the
given dimensions of the ladder and the application points of the forces we find:
−(4 ft)T sin θ + (8 ft)Nf0 sin φ = 0 (3.21)
And now we stop and rest and size up where we stand with the solution. We have found
five equations, namely Eqs. 3.17 through 3.21. Our unknowns are Nf , Nf0 , T , Fx and Fy .
So we have enough equations to solve the problem. The physics is done. Only the math
remains.
Here’s one way to go about solving this set of equations.
Notice that Fx only appears in 3.17, so we can wait until the very end to use it (if we
want to get Fx ). Combining 3.18 and 3.20 we get
Fy = 192 lb − Nf and Fy = Nf0
3.2. WORKED EXAMPLES 67

so
Nf0 = 192 lb − Nf
or
Nf = 192 lb − Nf0 (3.22)
From Eq. 3.21, we cancel a factor of 4 ft and find

2Nf0 sin φ
T = = 0.658Nf0 (3.23)
sin θ
Now we use results 3.22 and 3.23 in Eq. 3.19 to substitute for Nf and T . First, in 3.19
we can cancel a factor of 2 ft and rearrange to get:

−4Nf sin φ + 2T sin θ = −(192 lb) sin φ

Now substitute for Nf and T . This gives:

−4(192 lb − Nf0 ) sin φ + 2(0.658)Nf0 sin θ = −(192 lb) sin φ

Some more regrouping and evaluation of terms gives

(2.50)Nf0 = 180 lb

and at last we have an answer:


(180 lb)
Nf0 = = 72 lb
(2.50)

Having one solution, we can quickly get all the rest:

Nf = 192 lb − Nf0 = 120 lb

2(72 lb) sin φ


T = = 47 lb
sin θ
Glancing at what the problem asked us for, we see that we’ve now answered all the parts:

(a) T = 47 lb (b) Nf = 120 lb (c) Nf0 = 72 lb

But being the thorough kind of guy that I am, I’d like to find the components of the
force F which the ladder parts exert on each other at C. They are:

Fx = −T = −47 lb Fy = Nf0 = 72 lb

(So the direction that I chose for F rather arbitrarily in the figures was the correct one.)
68 CHAPTER 3. STATIC EQUILIBRIUM
Chapter 6

Work, Kinetic Energy and Potential


Energy

6.1 The Important Stuff


6.1.1 Kinetic Energy
For an object with mass m and speed v, the kinetic energy is defined as

K = 21 mv 2 (6.1)

Kinetic energy is a scalar (it has magnitude but no direction); it is always a positive
number; and it has SI units of kg · m2/ s2 . This new combination of the basic SI units is
known as the joule:
2
1 joule = 1 J = 1 kg·m
s2
(6.2)
As we will see, the joule is also the unit of work W and potential energy U. Other energy
units often seen are:
2
1 erg = 1 g·cm
s2
= 10−7 J 1 eV = 1.60 × 10−19 J

6.1.2 Work
When an object moves while a force is being exerted on it, then work is being done on the
object by the force.
If an object moves through a displacement d while a constant force F is acting on it, the
force does an amount of work equal to

W = F · d = F d cos φ (6.3)

where φ is the angle between d and F.


Work is also a scalar and has units of 1 N · m. But we can see that this is the same as
the joule, defined in Eq. 6.2.

127
128 CHAPTER 6. WORK, KINETIC ENERGY AND POTENTIAL ENERGY

Work can be negative; this happens when the angle between force and displacement is
larger than 90◦ . It can also be zero; this happens if φ = 90◦ . To do work, the force must
have a component along (or opposite to) the direction of the motion.
If several different (constant) forces act on a mass while it moves though a displacement
d, then we can talk about the net work done by the forces,

Wnet = F1 · d + F1 · d + F1 · d + . . . (6.4)
X 
= F ·d (6.5)
= Fnet · d (6.6)

If the force which acts on the object is not constant while the object moves then we must
perform an integral (a sum) to find the work done.
Suppose the object moves along a straight line (say, along the x axis, from xi to xf ) while
a force whose x component is Fx(x) acts on it. (That is, we know the force Fx as a function
of x.) Then the work done is Z xf
W = Fx(x) dx (6.7)
xi

Finally, we can give the most general expression for the work done by a force. If an object
moves from ri = xi i + yi j + zi k to rf = xf i + yf j + zf k while a force F(r) acts on it the work
done is: Z xf Z yf Z zf
W = Fx(r) dx + Fy (r) dy + Fz (r) dz (6.8)
xi yi zi

where the integrals are calculated along the path of the object’s motion. This expression
can be abbreviated as Z rf
W = F · dr . (6.9)
ri

This is rather abstract! But most of the problems where we need to calculate the work done
by a force will just involve Eqs. 6.3 or 6.7
We’re familiar with the force of gravity; gravity does work on objects which move ver-
tically. One can show that if the height of an object has changed by an amount ∆y then
gravity has done an amount of work equal to

Wgrav = −mg∆y (6.10)

regardless of the horizontal displacement. Note the minus sign here; if the object increases
in height it has moved oppositely to the force of gravity.

6.1.3 Spring Force


The most famous example of a force whose value depends on position is the spring force,
which describes the force exerted on an object by the end of an ideal spring. An ideal
spring will pull inward on the object attached to its end with a force proportional to the
amount by which it is stretched; it will push outward on the object attached to its with a
force proportional to amount by which it is compressed.
6.1. THE IMPORTANT STUFF 129

If we describe the motion of the end of the spring with the coordinate x and put the
origin of the x axis at the place where the spring exerts no force (the equilibrium position)
then the spring force is given by
Fx = −kx (6.11)
Here k is force constant, a number which is different for each ideal spring and is a measure
of its “stiffness”. It has units of N/ m = kg/ s2 . This equation is usually referred to as
Hooke’s law. It gives a decent description of the behavior of real springs, just as long as
they can oscillate about their equilibrium positions and they are not stretched by too much!
When we calculate the work done by a spring on the object attached to its end as the
object moves from xi to xf we get:

Wspring = 12 kx2i − 12 kx2f (6.12)

6.1.4 The Work–Kinetic Energy Theorem


One can show that as a particle moves from point ri to rf , the change in kinetic energy of
the object is equal to the net work done on it:

∆K = Kf − Ki = Wnet (6.13)

6.1.5 Power
In certain applications we are interested in the rate at which work is done by a force. If an
amount of work W is done in a time ∆t, then we say that the average power P due to the
force is
W
P = (6.14)
∆t
In the limit in which both W and ∆t are very small then we have the instantaneous power
P , written as:
dW
P = (6.15)
dt
The unit of power is the watt, defined by:
2
1 watt = 1 W = 1 Js = 1 kg·m
s3
(6.16)

The watt is related to a quaint old unit of power called the horsepower:

1 horsepower = 1 hp = 550 ft·lb


s
= 746 W

One can show that if a force F acts on a particle moving with velocity v then the
instantaneous rate at which work is being done on the particle is

P = F · v = F v cos φ (6.17)

where φ is the angle between the directions of F and v.


130 CHAPTER 6. WORK, KINETIC ENERGY AND POTENTIAL ENERGY

6.1.6 Conservative Forces


The work done on an object by the force of gravity does not depend on the path taken to
get from one position to another. The same is true for the spring force. In both cases we
just need to know the initial and final coordinates to be able to find W , the work done by
that force.
This situation also occurs with the general law for the force of gravity (Eq. 5.4.) as well
as with the electrical force which we learn about in the second semester!
This is a different situation from the friction forces studied in Chapter 5. Friction forces
do work on moving masses, but to figure out how much work, we need to know how the
mass got from one place to another.
If the net work done by a force does not depend on the path taken between two points,
we say that the force is a conservative force. For such forces it is also true that the net
work done on a particle moving on around any closed path is zero.

6.1.7 Potential Energy


For a conservative force it is possible to find a function of position called the potential
energy, which we will write as U(r), from which we can find the work done by the force.
Suppose a particle moves from ri to rf . Then the work done on the particle by a conser-
vative force is related to the corresponding potential energy function by:

Wri →rf = −∆U = U(ri ) − U(rf ) (6.18)

The potential energy U(r) also has units of joules in the SI system.
When our physics problems involve forces for which we can have a potential energy
function, we usually think about the change in potential energy of the objects rather than the
work done by these forces. However for non–conservative forces, we must directly calculate
their work (or else deduce it from the data given in our problems).

We have encountered two conservative forces so far in our study. The simplest is the
force of gravity near the surface of the earth, namely −mgj for a mass m, where the y axis
points upward. For this force we can show that the potential energy function is

Ugrav = mgy (6.19)

In using this equation, it is arbitrary where we put the origin of the y axis (i.e. what we call
“zero height”). But once we make the choice for the origin we must stick with it.
The other conservative force is the spring force. A spring of force constant k which is
extended from its equilibrium position by an amount x has a potential energy given by

Uspring = 21 kx2 (6.20)


6.1. THE IMPORTANT STUFF 131

6.1.8 Conservation of Mechanical Energy


If we separate the forces in the world into conservative and non-conservative forces, then the
work–kinetic energy theorem says

W = Wcons + Wnon−cons = ∆K

But from Eq. 6.18, the work done by conservative forces can be written as a change in
potential energy as:
Wcons = −∆U
where U is the sum of all types of potential energy. With this replacement, we find:

−∆U + Wnon−cons = ∆K

Rearranging this gives the general theorem of the Conservation of Mechanical Energy:

∆K + ∆U = Wnon−cons (6.21)

We define the total energy E of the system as the sum of the kinetic and potential
energies of all the objects:
E =K +U (6.22)
Then Eq. 6.21 can be written

∆E = ∆K + ∆U = Wnon−cons (6.23)

In words, this equation says that the total mechanical energy changes by the amount of work
done by the non–conservative forces.
Many of our physics problems are about situations where all the forces acting on the
moving objects are conservative; loosely speaking, this means that there is no friction, or
else there is negligible friction.
If so, then the work done by non–conservative forces is zero, and Eq. 6.23 takes on a
simpler form:
∆E = ∆K + ∆U = 0 (6.24)
We can write this equation as:

Ki + Ui = Kf + Uf or Ei = Ef

In other words, for those cases where we can ignore friction–type forces, if we add up all
the kinds of energy for the particle’s initial position, it is equal to the sum of all the kinds
of energy for the particle’s final position. In such a case, the amount of mechanical energy
stays the same. . . it is conserved.
Energy conservation is useful in problems where we only need to know about positions
or speeds but not time for the motion.
132 CHAPTER 6. WORK, KINETIC ENERGY AND POTENTIAL ENERGY

6.1.9 Work Done by Non–Conservative Forces


When the system does have friction forces then we must go back to Eq. 6.23. The change in
total mechanical energy equals the work done by the non–conservative forces:

∆E = Ef − Ei = Wnon−cons

(In the case of sliding friction with the rule fk = µk N it is possible to compute the work
done by the non–conservative force.)

6.1.10 Relationship Between Conservative Forces and Potential


Energy (Optional?)
Eqs. 6.9 (the general expression for work W ) and 6.18 give us a relation between the force F
on a particle (as a function of position, r) and the change in potential energy as the particle
moves from ri to rf : Zrf
F · dr = −∆U (6.25)
ri

Very loosely speaking, potential energy is the (negative) of the integral of F(r). Eq. 6.25
can be rewritten to show that (loosely speaking!) the force F(r) is the (minus) derivative of
U(r). More precisely, the components of F can be gotten by taking partial derivatives of U
with respect to the Cartesian coordinates:

∂U ∂U ∂U
Fx = − Fy = − Fz = − (6.26)
∂x ∂y ∂z

In case you haven’t come across partial derivatives in your mathematics education yet:
They come up when we have functions of several variables (like a function of x, y and z); if
we are taking a partial derivative with respect to x, we treat y and z as constants.
As you may have already learned, the three parts of Eq. 6.26 can be compactly written
as
F = −∇U
which can be expressed in words as “F is the negative gradient of U”.

6.1.11 Other Kinds of Energy


This chapter covers the mechanical energy of particles; later, we consider extended objects
which can rotate, and they will also have rotational kinetic energy. Real objects also have
temperature so that they have thermal energy. When we take into account all types of
energy we find that total energy is completely conserved. . . we never lose any! But here we
are counting only the mechanical energy and if (in real objects!) friction is present some of
it can be lost to become thermal energy.
6.2. WORKED EXAMPLES 133

6.2 Worked Examples


6.2.1 Kinetic Energy

1. If a Saturn V rocket with an Apollo spacecraft attached has a combined mass


of 2.9 × 105 kg and is to reach a speed of 11.2 km
s
, how much kinetic energy will it
then have? [HRW5 7-1]

(Convert some units first.) The speed of the rocket will be


103 m
!
v = (11.2 km
s
) = 1.12 × 104 m
s
.
1 km
We know its mass: m = 2.9 × 105 kg. Using the definition of kinetic energy, we have
K = 21 mv 2 = 12 (2.9 × 105 kg)(1.12 × 104 m 2
s
) = 1.8 × 1013 J
The rocket will have 1.8 × 1013 J of kinetic energy.

2. If an electron (mass m = 9.11 × 10−31 kg) in copper near the lowest possible
temperature has a kinetic energy of 6.7×10−19 J, what is the speed of the electron?
[HRW5 7-2]

Use the definition of kinetic energy, K = 21 mv 2 and the given values of K and m, and
solve for v. We find:
2K 2(6.7 × 10−19 J) m2
v2 = = = 1.47 × 1012 s2
m (9.11 × 10−31 kg)
which gives:
v = 1.21 × 106 m
s
m
The speed of the electron is 1.21 × 106 s
.

6.2.2 Work

3. A floating ice block is pushed through a displacement of d = (15 m)i − (12 m)j
along a straight embankment by rushing water, which exerts a force F = (210 N)i−
(150 N)j on the block. How much work does the force do on the block during the
displacement? [HRW5 7-11]

Here we have the simple case of a straight–line displacement d and a constant force F.
Then the work done by the force is W = F · d. We are given all the components, so we can
compute the dot product using the components of F and d:
W = F · d = Fxdx + Fy dy = (210 N)((15 m) + (−150 N)(−12 m) = 4950 J
134 CHAPTER 6. WORK, KINETIC ENERGY AND POTENTIAL ENERGY

4.00

3.00

Fx, N
2.00

1.00

0.00
0.0 5.0 10.0 15.0
x, m

Figure 6.1: Force Fx, which depends on position x; see Example 4.

The force does 4950 J of work.

4. A particle is subject to a force Fx that varies with position as in Fig. 6.1. Find
the work done by the force on the body as it moves (a) from x = 0 to x = 5.0 m,
(b) from x = 5.0 m to x = 10 m and (c) from x = 10 m to x = 15 m. (d) What is the
total work done by the force over the distance x = 0 to x = 15 m? [Ser4 7-23]

(a) Here the force is not the same all through the object’s motion, so we can’t use the
simple formula W = Fx x. We must use the more general expression for the work done when
a particle moves along a straight line,
Z xf
W = Fx dx .
xi

Of course, this is just the “area under the curve” of Fx vs. x from xi to xf .
In part (a) we want this “area” evaluated from x = 0 to x = 5.0 m. From the figure, we
see that this is just half of a rectangle of base 5.0 m and height 3.0 N. So the work done is
W = 12 (3.0 N)(5.0 m) = 7.5 J .
(Of course, when we evaluate the “area”, we just keep the units which go along with the
base and the height; here they were meters and newtons, the product of which is a joule.)
So the work done by the force for this displacement is 7.5 J.
(b) The region under the curve from x = 5.0 m to x = 10.0 m is a full rectangle of base
5.0 m and height 3.0 N. The work done for this movement of the particle is
W = (3.0 N)(5.0 m) = 15. J

(c) For the movement from x = 10.0 m to x = 15.0 m the region under the curve is a half
rectangle of base 5.0 m and height 3.0 N. The work done is
W = 12 (3.0 N)(5.0 m) = 7.5 J .
6.2. WORKED EXAMPLES 135

(d) The total work done over the distance x = 0 to x = 15.0 m is the sum of the three
separate “areas”,
Wtotal = 7.5 J + 15. J + 7.5 J = 30. J

5. What work is done by a force F = (2x N)i + (3 N)j, with x in meters, that moves
a particle from a position ri = (2 m)i + (3 m)j to a position rf = −(4 m)i − (3 m)j ?
[HRW5 7-31]

We use the general definition of work (for a two–dimensional problem),


Z xf Z yf
W = Fx (r) dx + Fy (r) dy
xi yi

With Fx = 2x and Fy = 3 [we mean that F in newtons when x is in meters; work W will
come out with units of joules!], we find:
Z −4 m Z −3 m
W = 2x dx + 3 dy
2m 3m
2 −4 m −3 m
= x + 3x
2m 3m
= [(16) − (4)] J + [(−9) − (9)] J
= −6 J

6.2.3 Spring Force

6. An archer pulls her bow string back 0.400 m by exerting a force that increases
from zero to 230 N. (a) What is the equivalent spring constant of the bow? (b)
How much work is done in pulling the bow? [Ser4 7-25]

(a) While a bow string is not literally spring, it may behave like one in that it exerts a force
on the thing attached to it (like a hand!) that is proportional to the distance of pull from
the equilibrium position. The correspondence is illustrated in Fig. 6.2.
We are told that when the string has been pulled back by 0.400 m, the string exerts a
restoring force of 230 N. The magnitude of the string’s force is equal to the force constant k
times the magnitude of the displacement; this gives us:

|Fstring| = 230 N = k(0.400 m)

Solving for k,
(230 N) N
k= = 575 m
(0.400 m)
N
The (equivalent) spring constant of the bow is 575 m .
136 CHAPTER 6. WORK, KINETIC ENERGY AND POTENTIAL ENERGY

F
F x
x
x=0
x=0
(a) (b)

Figure 6.2: The force of a bow string (a) on the object pulling it back can be modelled as as ideal spring
(b) exerting a restoring force on the mass attached to its end.

(b) Still treating the bow string as if it were an ideal spring, we note that in pulling the
string from a displacement of x = 0 to x = 0.400 m the string does as amount of work on
the hand given by Eq. 6.12:

Wstring = 1
2
kx2i − 21 kx2f
= 0 − 12 (575 mN
)(0.400 m)2
= −46.0 J

Is this answer to the question? Not quite. . . we were really asked for the work done by the
hand on the bow string. But at all times during the pulling, the hand exerted an equal and
opposite force on the string. The force had the opposite direction, so the work that it did
has the opposite sign. The work done (by the hand) in pulling the bow is +46.0 J.

6.2.4 The Work–Kinetic Energy Theorem

7. A 40 kg box initially at rest is pushed 5.0 m along a rough horizontal floor with
a constant applied horizontal force of 130 N. If the coefficient of friction between
the box and floor is 0.30, find (a) the work done by the applied force, (b) the
energy lost due to friction, (c) the change in kinetic energy of the box, and (d)
the final speed of the box. [Ser4 7-37]

(a) The motion of the box and the forces which do work on it are shown in Fig. 6.3(a). The
(constant) applied force points in the same direction as the displacement. Our formula for
the work done by a constant force gives

Wapp = F d cos φ = (130 N)(5.0 m) cos 0◦ = 6.5 × 102 J

The applied force does 6.5 × 102 J of work.


(b) Fig. 6.3(b) shows all the forces acting on the box.
6.2. WORKED EXAMPLES 137

d N

ffric ffric
Fapp Fapp

mg
(a) (b)

Figure 6.3: (a) Applied force and friction force both do work on the box. (b) Diagram showing all the
forces acting on the box.

The vertical forces acting on the box are gravity (mg, downward) and the floor’s normal
force (N, upward). It follows that N = mg and so the magnitude of the friction force is

ffric = µN = µmg = (0.30)(40 kg)(9.80 sm2 ) = 1.2 × 102 N

The friction force is directed opposite the direction of motion (φ = 180◦ ) and so the work
that it does is

Wfric = F d cos φ
= ffric d cos 180◦ = (1.2 × 102 N)(5.0 m)(−1) = −5.9 × 102 J

or we might say that 5.9 × 102 J is lost to friction.


(c) Since the normal force and gravity do no work on the box as it moves, the net work done
is
Wnet = Wapp + Wfric = 6.5 × 102 J − 5.9 × 102 J = 62 J .
By the work–Kinetic Energy Theorem, this is equal to the change in kinetic energy of the
box:
∆K = Kf − Ki = Wnet = 62 J .

(d) Here, the initial kinetic energy Ki was zero because the box was initially at rest. So we
have Kf = 62 J. From the definition of kinetic energy, K = 12 mv 2, we get the final speed of
the box:
2Kf 2(62 J) 2
vf2 = = = 3.1 ms2
m (40 kg)
so that
vf = 1.8 ms

8. A crate of mass 10.0 kg is pulled up a rough incline with an initial speed of


1.50 ms . The pulling force is 100 N parallel to the incline, which makes an angle of
20.0◦ with the horizontal. The coefficient of kinetic friction is 0.400, and the crate
138 CHAPTER 6. WORK, KINETIC ENERGY AND POTENTIAL ENERGY

0m
5.0 f = 110o
o
70

20o 20
o F

(a) (b)

Figure 6.4: (a) Block moves 5.00 m up plane while acted upon by gravity, friction and an applied force.
(b) Directions of the displacement and the force of gravity.

is pulled 5.00 m. (a) How much work is done by gravity? (b) How much energy
is lost due to friction? (c) How much work is done by the 100 N force? (d) What
is the change in kinetic energy of the crate? (e) What is the speed of the crate
after being pulled 5.00 m? [Ser4 7-47]

(a) We can calculate the work done by gravity in two ways. First, we can use the definition:
W = F · d. The magnitude of the gravity force is

Fgrav = mg = (10.0 kg(9.80 sm2 ) = 98.0 N

and the displacement has magnitude 5.00 m. We see from geometry (see Fig. 6.4(b)) that the
angle between the force and displacement vectors is 110◦ . Then the work done by gravity is

Wgrav = F d cos φ = (98.0 N)(5.00 m) cos 110◦ = −168 J .

Another way to work the problem is to plug the right values into Eq. 6.10. From simple
geometry we see that the change in height of the crate was

∆y = (5.00 m) sin 20◦ = +1.71 m

Then the work done by gravity was

Wgrav = −mg∆y = −(10.0 kg)(9.80 m


s2
)(1.71 m) = −168 J

(b) To find the work done by friction, we need to know the force of friction. The forces
on the block are shown in Fig. 6.5(a). As we have seen before, the normal force between
the slope and the block is mg cos θ (with θ = 20◦ ) so as to cancel the normal component of
the force of gravity. Then the force of kinetic friction on the block points down the slope
(opposite the motion) and has magnitude

fk = µk N = µmg cos θ
= (0.400)(10.0 kg)(9.80 sm2 ) cos 20◦ = 36.8 N
6.2. WORKED EXAMPLES 139

Fappl
fk

mg cos q
q = 20o 20o
mg

(a) (b)

Figure 6.5: (a) Gravity and friction forces which act on the block. (b) The applied force of 100 N is along
the direction of the motion.

This force points exactly opposite the direction of the displacement d, so the work done by
friction is
Wfric = fk d cos 180◦ = (36.8 N)(5.00 m)(−1) = −184 J
(c) The 100 N applied force pulls in the direction up the slope, which is along the direction
of the displacement d. So the work that is does is
Wappl = F d cos 0◦ = (100 N)(5.00 m)(1) = 500. J

(d) We have now found the work done by each of the forces acting on the crate as it moved:
Gravity, friction and the applied force. (We should note the the normal force of the surface
also acted on the crate, but being perpendicular to the motion, it did no work.) The net
work done was:
Wnet = Wgrav + Wfric + Wappl
= −168 J − 184 J + 500. J = 148 J
From the work–energy theorem, this is equal to the change in kinetic energy of the box:
∆K = Wnet = 148 J.
(e) The initial kinetic energy of the crate was
Ki = 1
2
(10.0 kg)(1.50 ms )2 = 11.2 J
If the final speed of the crate is v, then the change in kinetic energy was:
∆K = Kf − Ki = 21 mv 2 − 11.2 J .
Using our answer from part (d), we get:
2(159 J)
∆K = 12 mv 2 − 11.2 J = 148 J =⇒ v2 =
m
So then:
2(159 J) 2
v2 = = 31.8 ms2 =⇒ v = 5.64 ms .
(10.0 kg)
The final speed of the crate is 5.64 ms .
140 CHAPTER 6. WORK, KINETIC ENERGY AND POTENTIAL ENERGY

10.0 m
W= 700 N

Figure 6.6: Marine climbs rope in Example 9. You don’t like my drawing? Tell it to the Marines!

6.2.5 Power

9. A 700 N marine in basic training climbs a 10.0 m vertical rope at a constant


speed of 8.00 s. What is his power output? [Ser4 7-53]

Marine is shown in Fig. 6.6. The speed of the marine up the rope is
d 10.0 m
v= = = 1.25 ms
t 8.00 s
The forces acting on the marine are gravity (700 N, downward) and the force of the rope
which must be 700 N upward since he moves at constant velocity. Since he moves in the
same direction as the rope’s force, the rope does work on the marine at a rate equal to
dW
P = = F · v = F v = (700 N)(1.25 ms ) = 875 W .
dt
(It may be hard to think of a stationary rope doing work on anybody, but that is what is
happening here.)
This number represents a rate of change in the potential energy of the marine; the energy
comes from someplace. He is losing (chemical) energy at a rate of 875 W.

10. Water flows over a section of Niagara Falls at a rate of 1.2 × 106 kg/ s and falls
50 m. How many 60 W bulbs can be lit with this power? [Ser4 7-54]

Whoa! Waterfalls? Bulbs? What’s going on here??


If a certain mass m of water drops by a height h (that is, ∆y = −h), then from Eq. 6.10,
gravity does an amount of work equal to mgh. If this change in height occurs over a time
interval ∆t then the rate at which gravity does work is mgh/∆t.
For Niagara Falls, if we consider the amount of water that falls in one second, then a
mass m = 1.2 × 106 kg falls through 50 m and the work done by gravity is
Wgrav = mgh = (1.2 × 106 kg)(9.80 sm2 )(50 m) = 5.88 × 108 J .
6.2. WORKED EXAMPLES 141

v
vy=0 (Max height)

v0= 20 m/s h

34o

Figure 6.7: Snowball is launched at angle of 34◦ in Example 11.

This occurs every second, so gravity does work at a rate of

mgh 5.88 × 108 J


Pgrav = = = 5.88 × 108 W
∆t 1s
As we see later, this is also the rate at which the water loses potential energy. This energy
can be converted to other forms, such as the electrical energy to make a light bulb function.
In this highly idealistic example, all of the energy is converted to electrical energy.
A 60 W light bulb uses energy at a rate of 60 Js = 60 W. We see that Niagara Falls
puts out energy at a rate much bigger than this! Assuming all of it goes to the bulbs, then
dividing the total energy consumption rate by the rate for one bulb tells us that

5.88 × 108 W
N= = 9.8 × 106
60 W
bulbs can be lit.

6.2.6 Conservation of Mechanical Energy

11. A 1.50 kg snowball is shot upward at an angle of 34.0◦ to the horizontal with
an initial speed of 20.0 ms . (a) What is its initial kinetic energy? (b) By how much
does the gravitational potential energy of the snowball–Earth system change as
the snowball moves from the launch point to the point of maximum height? (c)
What is that maximum height? [HRW5 8-31]

(a) Since the initial speed of the snowball is 20.0 ms , we have its initial kinetic energy:

Ki = 12 mv02 = 12 (1.50 kg)(20.0 ms )2 = 300. J

(b) We need to remember that since this projectile was not fired straight up, it will still
have some kinetic energy when it gets to maximum height! That means we have to think a
little harder before applying energy principles to answer this question.
142 CHAPTER 6. WORK, KINETIC ENERGY AND POTENTIAL ENERGY

At maximum height, we know that the y component of the snowball’s velocity is zero.
The x component is not zero.
But we do know that since a projectile has no horizontal acceleration, the x component
will remain constant; it will keep its initial value of

v0x = v0 cos θ0 = (20.0 ms ) cos 34◦ = 16.6 ms

so the speed of the snowball at maximum height is 16.6 ms . At maximum height, (the final
position) the kinetic energy is

Kf = 12 mvf2 = 12 (1.50 kg)(16.6 ms )2 = 206. J

In this problem there are only conservative forces (namely, gravity). The mechanical
energy is conserved:
Ki + Ui = Kf + Uf
We already found the initial kinetic energy of the snowball: Ki = 300. J. Using Ugrav = mgy
(with y = 0 at ground level), the initial potential energy is Ui = 0. Then we can find the
final potential energy of the snowball:

Uf = Ki + Ui − Kf
= 300. J + 0 − 206. J
= 94. J

The final gravitational potential energy of the snowball–earth system (a long–winded way
of saying what U is!) is then 94. J. (Since its original value was zero, this is the answer to
part (b).)
(c) If we call the maximum height of the snowball h, then we have

Uf = mgh

Solve for h:
Uf (94. J)
h= = = 6.38 m
mg (1.5 kg)(9.80 sm2 )
The maximum height of the snowball is 6.38 m.

12. A pendulum consists of a 2.0 kg stone on a 4.0 m string of negligible mass. The
stone has a speed of 8.0 ms when it passes its lowest point. (a) What is the speed
when the string is at 60◦ to the vertical? (b) What is the greatest angle with the
vertical that the string will reach during the stone’s motion? (c) If the potential
energy of the pendulum–Earth system is taken to be zero at the stone’s lowest
point, what is the total mechanical energy of the system? [HRW5 8-32]

(a) The condition of the pendulum when the stone passes the lowest point is shown in
Fig. 6.8(a). Throughout the problem we will measure the height y of the stone from the
6.2. WORKED EXAMPLES 143

60o

4.0 m

v= ?
2.0 kg

8.0 m/s

(a) (b)

Figure 6.8: (a) Pendulum in Example 12 swings through lowest point. (b) Pendulum has swung 60◦ past
lowest point.

bottom of its swing. Then at the bottom of the swing the stone has zero potential energy,
while its kinetic energy is

Ki = 12 mv02 = 12 (2.0 kg)(8.0 ms )2 = 64 J

When the stone has swung up by 60◦ (as in Fig. 6.8(b)) it has some potential energy. To
figure out how much, we need to calculate the height of the stone above the lowest point of
the swing. By simple geometry, the stone’s position is

(4.0 m) cos 60◦ = 2.0 m

down from the top of the string, so it must be

4.0 m − 2.0 m = 2.0 m

up from the lowest point. So its potential energy at this point is

Uf = mgy = (2.0 kg)(9.80 sm2 )(2.0 m) = 39.2 J

It will also have a kinetic energy Kf = 21 mvf2 , where vf is the final speed.
Now in this system there are only a conservative force acting on the particle of interest,
i.e. the stone. (We should note that the string tension also acts on the stone, but since
it always pulls perpendicularly to the motion of the stone, it does no work.) So the total
mechanical energy of the stone is conserved:

Ki + Ui = Kf + Uf

We can substitute the values found above to get:

64.0 J + 0 = 21 (2.0 kg)vf2 + 39.2 J

which we can solve for vf :


2
(1.0 kg)vf2 = 64.0 J − 39.2 J = 24.8 J =⇒ vf2 = 24.8 ms2
144 CHAPTER 6. WORK, KINETIC ENERGY AND POTENTIAL ENERGY

4.0 m

4.0 m

Figure 6.9: Stone reaches its highest position in the swing, which we specify by some angle θ measured
above the horizontal.

and then:
vf = 5.0 ms
The speed of the stone at the 60◦ position will be 5.0 ms .
(b) Clearly, since the stone is still in motion at an angle of 60◦ , it will keep moving to greater
angles and larger heights above the bottom position. For all we know, it may keep rising
until it gets to some angle θ above the position where the string is horizontal, as shown
in Fig. 6.9. We do assume that the string will stay straight until this point, but that is a
reasonable assumption.
Now at this point of maximum height, the speed of the mass is instantaneously zero. So
in this final position, the kinetic energy is Kf = 0. Its height above the starting position is

y = 4.0 m + (4.0 m) sin θ = (4.0 m)(1 + sin θ) (6.27)

so that its potential energy there is

Uf = mgyf = (2.0 kg)(9.80 sm2 )(4.0 m)(1 + sin θ) = (78.4 J)(1 + sin θ)

We use the conservation of mechanical energy (from the position at the bottom of the swing)
to find θ: Ki + Ui = Kf + Uf , so:

Uf = Ki + Ui − Kf =⇒ (78.4 J)(1 + sin θ) = 64 J + 0 − 0

This gives us:


78.4 J
1 + sin θ = = 1.225 =⇒ sin θ = 0.225
64 J
and finally
θ = 13◦
We do get a sensible answer of θ so we were right in writing down Eq. 6.27. Actually this
equation would also have been correct if θ were negative and the pendulum reached its
highest point with the string below the horizontal.
6.2. WORKED EXAMPLES 145

h
R

Figure 6.10: Bead slides on track in Example 13.

13. A bead slides without friction on a loop–the–loop track (see Fig. 6.10). If
the bead is released from a height h = 3.50R, what is its speed at point A? How
large is the normal force on it if its mass is 5.00 g? [Ser4 8-11]

In this problem, there are no friction forces acting on the particle (the bead). Gravity acts
on it and gravity is a conservative force. The track will exert a normal forces on the bead,
but this force does no work. So the total energy of the bead —kinetic plus (gravitational)
potential energy— will be conserved.
At the initial position, when the bead is released, the bead has no speed; Ki = 0. But
it is at a height h above the bottom of the track. If we agree to measure height from the
bottom of the track, then the initial potential energy of the bead is

Ui = mgh

where m = 5.00 g is the mass of the bead.


At the final position (A), the bead has both kinetic and potential energy. If the bead’s
speed at A is v, then its final kinetic energy is Kf = 12 mv 2. At position A its height is 2R
(it is a full diameter above the “ground level” of the track) so its potential energy is

Uf = mg(2R) = 2mgR .

The total energy of the bead is conserved: Ki + Ui = Kf + Uf . This gives us:

0 + mgh = 12 mv 2 + 2mgR ,

where we want to solve for v (the speed at A). The mass m cancels out, giving:

gh = 21 v 2 = 2gR =⇒ 1 2
2
v = gh − 2gR = g(h − 2R)

and then
146 CHAPTER 6. WORK, KINETIC ENERGY AND POTENTIAL ENERGY

N
mg
R
Dir. of
accel.

Figure 6.11: Forces acting on the bead when it is at point A (the top of the loop).

v 2 = 2g(h − 2R) = 2g(3.50R − 2R) = 2g(1.5R) = 3.0 gR (6.28)


and finally q
v= 3.0 gR .
Since we don’t have a numerical value for R, that’s as far as we can go.
In the next part of the problem, we think about the forces acting on the bead at point A.
These are diagrammed in Fig. 6.11. Gravity pulls down on the bead with a force mg. There
is also a normal force from the track which I have drawn as having a downward component
N. But it is possible for the track to be pushing upward on the bead; if we get a negative
value for N we’ll know that the track was pushing up.
At the top of the track the bead is moving on a circular path of radius R, with speed v.
So it is accelerating toward the center of the circle, namely downward. We know that the
downward forces must add up to give the centripetal force mv 2/R:

mv 2 mv 2 v2
!
mg + N = =⇒ n= − mg = m −g .
R R R

But we can use our result from Eq. 6.28 to substitute for v 2. This gives:
3.0 gR
 
N =m − g = m(2g) = 2mg
R
Plug in the numbers:

N = 2(5.00 × 10−3 kg)(9.80 sm2 ) = 9.80 × 10−2 N

At point A the track is pushing downward with a force of 9.80 × 10−2 N.

14. Two children are playing a game in which they try to hit a small box on the
floor with a marble fired from a spring –loaded gun that is mounted on the table.
The target box is 2.20 m horizontally from the edge of the table; see Fig. 6.12.
6.2. WORKED EXAMPLES 147

2.20 m

Figure 6.12: Spring propels marble off table and hits (or misses) box on the floor.

1.10 cm
v

(a) (b)
Figure 6.13: Marble propelled by the spring–gun: (a) Spring is compressed, and system has potential
energy. (b) Spring is released and system has kinetic energy of the marble.

Bobby compresses the spring 1.10 cm, but the center of the marble falls 27.0 cm
short of the center of the box. How far should Rhoda compress the spring to
score a direct hit? [HRW5 8-36]

Let’s put the origin of our coordinate system (for the motion of the marble) at the edge
of the table. With this choice of coordinates, the object of the game is to insure that the x
coordinate of the marble is 2.20 m when it reaches the level of the floor.
There are many things we are not told in this problem! We don’t know the spring
constant for the gun, or the mass of the marble. We don’t know the height of the table
above the floor, either!
When the gun propels the marble, the spring is initially compressed and the marble is
motionless (see Fig. 6.13(a).) The energy of the system here is the energy stored in the spring,
Ei = 21 kx2 , where k is the force constant of the spring and x is the amount of compression
of the spring.) When the spring has returned to its natural length and has given the marble
a speed v, then the energy of the system is Ef = 21 mv 2. If we can neglect friction then
mechanical energy is conserved during the firing, so that Ef = Ei , which gives us:
s s
k 2 k
1
2
mv 2 = 1
2
kx2 =⇒ v= x =x
m m
148 CHAPTER 6. WORK, KINETIC ENERGY AND POTENTIAL ENERGY

We will let x and v be the compression and initial marble speed for Bobby’s attempt. Then
we have: s
k
v = (1.10 × 10−2 m) (6.29)
m
The marble’s trip from the edge of the table to the floor is (by now!) a fairly simple
kinematics problem. If the time the marble spends in the air is t and the height of the table
is h then the equation for vertical motion tells us:
h = 12 gt2 .
(This is true because the marble’s initial velocity is all horizontal. We do know that on
Bobby’s try, the marble’s x coordinate at impact was
x = 2.20 m − 0.27 m = 1.93 m
and since the horizontal velocity of the marble is v, we have:
vt = 1.93 m . (6.30)
There are too many unknowns to solve for k, v, h and t. . . but let’s go on.
Let’s suppose that Rhoda compresses the spring by an amount x0 so that the marble is
given a speed v 0. As before, we have
1 2 2
2
mv 0 = 12 kx0
(it’s the same spring and marble so that k and m are the same) and this gives:
s
k
v 0 = x0
. (6.31)
m
Now when Rhoda’s shot goes off the table and through the air, then if its time of flight is t0
then the equation for vertical motion gives us:
2
h = 21 gt0 .
This is the same equation as for t, so that the times of flight for both shots is the same:
t0 = t. Since the x coordinate of the marble for Rhoda’s shot will be x = 2.20 m, the equation
for horizontal motion gives us
v 0t = 2.20 m (6.32)
What can we do with these equations? If we divide Eq. 6.32 by Eq. 6.30 we get:
v 0t v0 2.20
= = = 1.14
vt v 1.93
If we divide Eq. 6.31 by Eq. 6.29 we get:
q
k
v 0 x0m x0
= q = .
v k
(1.10 × 10−2 m) m (1.10 × 10−2 m)
With these last two results, we can solve for x0 . Combining these equations gives:
x0
1.14 = =⇒ x0 = 1.14(1.10 × 10−2 m) = 1.25 cm
(1.10 × 10−2 m)
Rhoda should compress the spring by 1.25 cm in order to score a direct hit.
6.2. WORKED EXAMPLES 149

mk = 0.40
m= 3.0 kg

M= 5.0 kg

Figure 6.14: Moving masses in Example 15. There is friction between the surface and the 3.0 kg mass.

N d

fk

T
d

mg
(a) (b)

Figure 6.15: (a) Forces acting on m. (b) Masses m and M travel a distance d = 1.5 m as they increase in
speed from 0 to v.

6.2.7 Work Done by Non–Conservative Forces

15. The coefficient of friction between the 3.0 kg mass and surface in Fig. 6.14 is
0.40. The system starts from rest. What is the speed of the 5.0 kg mass when it
has fallen 1.5 m? [Ser4 8-25]

When the system starts to move, both masses accelerate; because the masses are con-
nected by a string, they always have the same speed . The block (m) slides on the rough
surface, and friction does work on it. Since its height does not change, its potential energy
does not change, but its kinetic energy increases. The hanging mass (M) drops freely; its
potential energy decreases but its kinetic energy increases.
We want to use energy principles to work this problem; since there is friction present,
we need to calculate the work done by friction.
The forces acting on m are shown in Fig. 6.15(a). The normal force N must be equal to
mg, so the force of kinetic friction on m has magnitude µk N = µk mg. This force opposes
150 CHAPTER 6. WORK, KINETIC ENERGY AND POTENTIAL ENERGY

the motion as m moves a distance d = 1.5 m, so the work done by friction is

Wfric = fk d cos φ = (µk mg)(d)(−1) = −(0.40)(3.0 kg)(9.80 sm2 )(1.5 m) = −17.6 J .

Mass m’s initial speed is zero, and its final speed is v. So its change in kinetic energy is

∆K = 12 (3.0 kg)v 2 − 0 = (1.5 kg)v 2

As we noted, m has no change in potential energy during the motion.


Mass M’s change in kinetic energy is

∆K = 12 (5.0 kg)v 2 − 0 = (2.5 kg)v 2

and since it has a change in height given by −d, its change in (gravitational) potential energy
is
∆U = Mg∆y = (5.0 kg)(9.80 sm2 )(−1.5 m) = −73.5 J
Adding up the changes for both masses, the total change in mechanical energy of this system
is

∆E = (1.5 kg)v 2 + (2.5 kg)v 2 − 73.5 J


= (4.0 kg)v 2 − 73.5 J

Now use ∆E = Wfric and get:

(4.0 kg)v 2 − 73.5 J = −17.6 J

Solve for v:
55.9 J 2
(4.0 kg)v 2 = 55.9 J =⇒ v2 = = 14.0 ms2
4.0 kg
which gives
v = 3.74 ms
The final speed of the 5.0 kg mass (in fact of both masses) is 3.74 ms .

16. A 10.0 kg block is released from point A in Fig. 6.16. The track is friction-
less except for the portion BC, of length 6.00 m. The block travels down the
track, hits a spring of force constant k = 2250 N/ m, and compresses it 0.300 m
from its equilibrium position before coming to rest momentarily. Determine the
coefficient of kinetic friction between surface BC and block. [Ser4 8-35]

We know that we must use energy methods to solve this problem, since the path of the
sliding mass is curvy.
The forces which act on the mass as it descends and goes on to squish the spring are:
gravity, the spring force and the force of kinetic friction as it slides over the rough part.
Gravity and the spring force are conservative forces, so we will keep track of them with the
potential energy associated with these forces. Friction is a non-conservative force, but in
6.2. WORKED EXAMPLES 151

A 10.0 kg

3.00 m
k
B C

6.00 m

Figure 6.16: System for Example 16

0.300 m

Equil. pos.
of spring v=0

Figure 6.17: After sliding down the slope and going over the rough part, the mass has maximally squished
the spring by an amount x = 0.300 m.

this case we can calculate the work that it does. Then, we can use the energy conservation
principle,
∆K + ∆U = Wnon−cons (6.33)
to find the unknown quantity in this problem, namely µk for the rough surface. We can
get the answer from this equation because we have numbers for all the quantities except for
Wnon−cons = Wfriction which depends on the coefficient of friction.
The block is released at point A so its initial speed (and hence, kinetic energy) is zero:
Ki = 0. If we measure height upwards from the level part of the track, then the initial
potential energy for the mass (all of it gravitational) is

Ui = mgh = (10.0 kg)(9.80 sm2 )(3.00 m) = 2.94 × 102 J

Next, for the “final” position of the mass, consider the time at which it has maximally
compressed the spring and it is (instantaneously) at rest. (This is shown in Fig. 6.17.) We
don’t need to think about what the mass was doing in between these two points; we don’t
care about the speed of the mass during its slide.
At this final point, the mass is again at rest, so its kinetic energy is zero: Kf = 0. Being
at zero height, it has no gravitational potential energy but now since there is a compressed
152 CHAPTER 6. WORK, KINETIC ENERGY AND POTENTIAL ENERGY

spring, there is stored (potential) energy in the spring. This energy is given by:

Uspring = 21 kx2 = 12 (2250 m


N
)(0.300 m)2 = 1.01 × 102 J

so the final potential energy of the system is Uf = 1.01 × 102 J.


The total mechanical energy of the system changes because there is a non–conservative
force (friction) which does work. As the mass (m) slides over the rough part, the vertical
forces are gravity (mg, downward) and the upward normal force of the surface, N. As there
is no vertical motion, N = mg. The magnitude of the force of kinetic friction is

fk = µk N = µk mg = µk (10.0 kg)(9.80 sm2 ) = µk (98.0 N)

As the block moves 6.00 m this force points opposite (180◦ from ) the direction of motion.
So the work done by friction is

Wfric = fk d cos φ = µk (98.0 N)(6.00 m) cos 180◦ = −µk (5.88 × 102 J)

We now have everything we need to substitute into the energy balance condition, Eq. 6.33.
We get:
(0 − 0) + (1.01 × 102 J − 2.94 × 102 J) = −µk (5.88 × 102 J) .
The physics is done. We do algebra to solve for µk :

−1.93 × 102 J = −µk (5.88 × 102 J) =⇒ µk = 0.328

The coefficient of kinetic friction for the rough surface and block is 0.328.

6.2.8 Relationship Between Conservative Forces and Potential En-


ergy (Optional?)

17. A potential energy function for a two–dimensional force is of the form U =


3x3 y − 7x. Find the force that acts at the point (x, y). [Ser4 8-39]

(We presume that the expression for U will give us U in joules when x and y are in
meters!)
We use Eq. 6.26 to get Fx and Fy :

∂U
Fx = −
∂x

= − (3x3 y − 7x)
∂x
= −(9x2 y − 7) = −9x2 y + 7
6.2. WORKED EXAMPLES 153

and:
∂U
Fy = −
∂y

= − (3x3 y − 7x)
∂y
= −(3x3 ) = −3x3

Then in unit vector form, F is:

F = (−9x2y + 7)i + (−3x3)j

where, if x and y are in meters then F is in newtons. Got to watch those units!
154 CHAPTER 6. WORK, KINETIC ENERGY AND POTENTIAL ENERGY
Chapter 4

Oscillatory Motion

4.1 The Important Stuff


4.1.1 Simple Harmonic Motion
In this chapter we consider systems which have a motion which repeats itself in time, that is,
it is periodic. In particular we look at systems which have some coordinate (say, x) which
has a sinusoidal dependence on time. A graph of x vs. t for this kind of motion is shown in
Fig. 4.1. Suppose a particle has a periodic, sinusoidal motion on the x axis, and its motion
takes it between x = +A and x = −A. Then the general expression for x(t) is

x(t) = A cos(ωt + φ) (4.1)

A is called the amplitude of the motion. For reasons which will become clearer later, ω is
called the angular frequency. We say that a mass which has a motion of the type given
in Eq. 4.1 undergoes simple harmonic motion.
From 4.1 we see that when the time t increases by an amount 2π ω
, the argument of the
cosine increases by 2π and the value of x will be the same. So the motion repeats itself
after a time interval 2π
ω
, which we denote as T , the period of the motion. The number of

Figure 4.1: Plot of x vs. t for simple harmonic motion. (t and x axes are unspecified!)

69
70 CHAPTER 4. OSCILLATORY MOTION

oscillations per time is given by f = T1 , called the frequency of the motion:

2π 1 ω
T = f= = (4.2)
ω T 2π

Rearranging we have a formula for ω in terms of f or T :


ω = 2πf = (4.3)
T
Though ω (angular frequency) and f (frequency) are closely related (with just a factor of
2π between them, we need to be careful to distinguish them; to help in this, we normally
express ω in units of rad
s
and f in units of cycle
s
, or Hz (Hertz). However, the real dimensions
of both are 1s in the SI system.
From x(t) we get the velocity of the particle:

dx
v(t) = = −ωA sin(ωt + φ) (4.4)
dt
and its acceleration:
dv
a(t) = = −ω 2A cos(ωt + φ) (4.5)
dt
We note that the maximum values of v and a are:

vmax = ωA amax = ω 2 A (4.6)

The maximum speed occurs in the middle of the oscillation. (The slope of x vs. t is greatest
in size when x = 0.) The magnitude of the acceleration is greatest at the ends of the
oscillation (when x = ±A).
Comparing Eq. 4.5 and Eq. 4.1 we see that

d2 x
= −ω 2 x (4.7)
dt2

which is the same as a(t) = −ω 2x(t). Using 4.1 and 4.4 and some trig we can also arrive at
a relation between the speed |v(t)| of the mass and its coordinate x(t):
q
|v(t)| = ωA| sin(ωt + φ)| = ωA 1 − cos2 (ωt + φ)
v
u !2
u x(t)
= ωA 1 −
t
. (4.8)
A

We could also arrive at this relation using energy conservation (as discussed below). Note,
if we are given x we can only give the absolute value of v since there are two possibilities for
velocity at each x (namely a ± pair).
4.1. THE IMPORTANT STUFF 71

k x

Figure 4.2: Mass m is attached to horizontal spring of force constant k; it slides on a frictionless surface!

4.1.2 Mass Attached to a Spring


Suppose a mass m is attached to the end of a spring of force constant k (whose other end is
fixed) and slides on a frictionless surface. This system is illustrated in Fig. 4.2. Then if we
measure the coordinate x of the mass from the place where it would be if the spring were at
its equilibrium length, Newton’s 2nd law gives

d2 x
Fx = −kx = max = m ,
dt2
and then we have
d2 x k
= − x. (4.9)
dt2 m
Comparing Eqs. 4.9 and 4.7 we can identify ω 2 with k
m
so that
s
k
ω= (4.10)
m
From the angular frequency ω we can find the period T and frequency f of the motion:
s
2π m 1 1 k
r
T = = 2π f= = (4.11)
ω k T 2π m
It should be noted that ω (and hence T and f) does not depend on the amplitude A
of the motion of the mass. In reality, of course if the motion of the mass is too large then
then spring will not obey Hooke’s Law so well, but as long as the oscillations are “small”
the period is the same for all amplitudes.

In the lab, it’s much easier to work with a mass bobbing up and down on a vertical
spring. One can (and should!) ask if we can still use the same formulae for T and f, or if
gravity (g) enters in somehow. In fact, the same formulae (Eq. 4.11) do apply in this case.
To be more clear about the vertical mass–spring system, we show such a system in
Fig. 4.3. In (a), the spring is oriented vertically and has some unstretched length. (We are
ignoring the mass of the spring.) When a mass m is attached to the end, the system will be
72 CHAPTER 4. OSCILLATORY MOTION

x
m
m

(a) (b) (c)

Figure 4.3: (a) Unstretched vertical spring of force constant k (assumed massless). (b) Mass attached to
spring is at equilibrium when the spring has been extended by a distance mg/k. (c) Mass will undergo small
oscillations about the new equilibrium position.

at equilibrium when the spring has been extended by some length y; balancing forces on the
mass, this extension is given by:
mg
ky = mg =⇒ y= .
k
When the mass is disturbed from its equilibrium position, it will undergo harmonic oscil-
lations which can be described by some coordinate x, where x is measured from the new
equilibrium position of the end of the spring. Then the motion is just like that of the
horizontal spring.

Finally, we note that for more precise work with a real spring–mass system one does need
to take into account the mass of the spring. If the spring has a total mass ms , one can show
that Eq. 4.10 should be modified to:
v
k
u
u
ω= t
ms (4.12)
m+ 3

That is, we replace the value of the mass m by m plus one–third the spring’s mass.

4.1.3 Energy and the Simple Harmonic Oscillator


For the mass–spring system, the kinetic energy is given by
K = 12 mv 2 = 21 mω 2A2 sin2 (ωt + φ) (4.13)
and the potential energy is

U = 12 kx2 = 12 kA2 cos2(ωt + φ) . (4.14)


k
Using ω 2 = m
in 4.13 we then find that the total energy is
E = K + U = 21 kA2 [sin2 (ωt + φ) + cos2 (ωt + φ)]
4.1. THE IMPORTANT STUFF 73

and the trig identity sin2 θ + cos2 θ = 1 gives

E = 21 kA2 (4.15)

showing that the energy of the simple harmonic oscillator (as typified by a mass on a spring)
is constant and is equal to the potential energy of the spring when it is maximally extended
(at which time the mass is motionless).

It is useful to use the principle of energy conservation to derive some general relations for
1–dimensional harmonic motion. (We will not use the particular parameters for the mass–
spring system, just the quantities contained in Eq. 4.1, which describes the motion of a mass
m along the x axis. From Eq. 4.13 we have the kinetic energy as a function of time

K = 12 mv 2 = 21 mω 2A2 sin2 (ωt + φ)

Now the maximum value of the kinetic energy is 12 mω 2 A2, which occurs when x = 0. Since
we are free to fix the “zero–point” of the potential energy, we can agree that U(x) = 0 at
x = 0. Then the total energy of the system must be equal to the maximum (i.e. x = 0 value
of the kinetic energy:
E = 12 mω 2A2
Then using these expressions, the potential energy of the system is

U = E−K
= 21 mω 2 A2 − 21 mω 2A2 sin2(ωt + φ) = 21 mω 2 A2(1 − sin2 (ωt + φ))
= 12 mω 2 A2 cos2 (ωt + φ)
= 12 mω 2 x2

Of course, for the mass–spring system U is given by 21 kx2, which gives the relation mω 2 = k,
q
k
or ω = m , which we’ve already found. If we use the relation vmax = ωA then the potential
energy can be written as
2
mvmax
U(x) = 12 mω 2 x2 = 12 2
x2 (4.16)
A

4.1.4 Relation to Uniform Circular Motion


There is a correspondence between simple harmonic motion and uniform circular motion,
which is illustrated in Fig. 4.4 (a) and (b). In (a) a mass point moves in a horizontal circular
path with uniform circular motion at a radius R (for example, it might be glued to the edge
of a spinning disk of radius R). Its angular velocity is ω, so its location is given by the
time–varying angle θ, where
θ(t) = ωt + φ
.
74 CHAPTER 4. OSCILLATORY MOTION

R
q(t)
x x
-R 0 +R

(a) (b)

Figure 4.4: (a) Mass point moves in a horizontal circle of radius R. The angular velocity of its motion is
ω. A guy with a big nose (seen from above) is observing the motion of the mass at the level of the circle.
He sees only the x coordinate of the point’s motion. (b) Motion of the mass as seen by the guy with the big
nose. The projection of the motion is the same as simple harmonic motion with angular frequency ω and
amplitude R.

Pivot I, M
q L d
q
CM
m

(a) (b)

Figure 4.5: (a) Simple pendulum. (b) Physical pendulum.

In 4.4 (b) we show the motion of the mass as it would be seen by someone looking
toward the +y direction at the level of the disk. Such an observer sees only the changing x
coordinate of the mass’s motion. Since x = R cos θ, the observed coordinate is

x(t) = R cos(θ(t)) = R cos(ωt + φ) ,

the same as Eq. 4.1. The motion of the corresponding (projected) harmonic oscillator has
an angular frequency of ω and an amplitude of R.

4.1.5 The Pendulum


We start with the simple pendulum, which has just a small mass m hanging from a string
of length L whose mass we can ignore. (See Fig. 4.5 (a).) The mass is set into motion so
4.1. THE IMPORTANT STUFF 75

that it moves in a vertical plane. One can show that if θ is the angle which the string makes
with the vertical, it obeys the differential equation:
d2 θ g
= − sin θ
dt2 L
One should note that this is not of the form given in Eq. 4.7.
Things are much simpler when we restrict θ to be “small” at all times. If that is the case,
then we can use the approximation sin θ ≈ θ, which is true if we are measuring θ in radians.
Then the differential equation becomes
d2 θ g
2
=− θ (4.17)
dt L
Comparison of this equation with Eq. 4.7 lets us identify the angular frequency of the
motion:
g
r
ω= (4.18)
L
s
2π L ω 1 g
r
T = = 2π f= = (4.19)
ω g 2π 2π L
The (perhaps) surprising thing about Eqs. 4.18 and 4.19 is that they have no dependence
on the mass suspended from the string or on the amplitude of the swing. . . as long as it is a
small angle!

θ(t) = θmax cos(ωt + φ) (4.20)


We must always keep our assumption of “small” θ in the back of our minds whenever we
do a problem with a pendulum. The formulae giving T and f become less accurate as θmax
gets too big.
An important generalization of the simple pendulum is that of a rigid body which is free
to rotate in a plane about some (frictionless!) pivot. Such a system is known as a physical
pendulum and is diagrammed in Fig. 4.5 (b).
Suppose we look at the line which joins the pivot to the center of mass of the object. If
θ is the angle which this line makes with the vertical, and if we again use the approximation
sin θ ≈ θ, one can show that it obeys the differential equation
d2 θ Mgd
2
=− θ (4.21)
dt I
where d is the distance between the pivot and the center of mass, M is the mass of the
object and I is the moment of inertia of the object about the given axis. (Note: the axis is
probably not at the center of mass; if it were, the mass wouldn’t oscillate!)
Following the usual procedure we find the period T :
s
2π I
T = = 2π (4.22)
ω Mgd
76 CHAPTER 4. OSCILLATORY MOTION

4.2 Worked Examples


4.2.1 Simple Harmonic Motion

1. The displacement of a particle at t = 0.25 s is given by the expression x =


(4.0 m) cos(3.0πt + π) where x is in meters and t is in seconds. Determine (a) the
frequency and period of the motion, (b) the amplitude of the motion, (c) the
phase constant, and (d) the displacement of the particle at t = 0.25 s. [Ser4 13-1]

(a) We compare the given function x(t) with the standard form for simple harmonic motion
given in Eq. 4.1. This gives us the angular frequency ω:
rad
ω = 3.0π s

and from this we can get the frequency and period:


ω 3.0π rad
s
f= = = 1.50 Hz
2π 2π
1 1
T = = = 0.667 s
f (1.50 s−1 )
(b) We easily read off the amplitude as the factor (a length) which multiplies the cosine
function:
A = 4.0 m

(c) Again, comparison with Eq. 4.1 gives


φ=π

(d) At t = 0.25 s the displacement (i.e. the coordinate) of the particle is:
x(0.25 s) = (4.0 m) cos((3.0π)(0.25) + π) = (4.0 m) cos((1.75)π)
= (4.0 m)(0.707) = 2.83 m

2. A loudspeaker produces a musical sound by means of the oscillation of a


diaphragm. If the amplitude of oscillation is limited to 1.0 × 10−3 mm, what fre-
quencies will result in the magnitude of the diaphragm’s acceleration exceeding
g? [HRW5 16-5]

We are given the amplitude of the diaphragm’s motion, A = 1.0×10−3 mm = 1.0×10−6 m.


From Eq. 4.6, the maximum value of the acceleration is amax = Aω 2. So then the angular
frequency that results in a maximum acceleration of g is
amax (9.8 sm2 )
ω2 = = = 9.8 × 106 s−2
A (1.0 × 10−6 m)
4.2. WORKED EXAMPLES 77

=⇒ ω = 3.1 × 103 s−1 .


This corresponds to a frequency of
ω (3.1 × 103 s−1 )
f= = = 5.0 × 102 Hz
2π 2π
At frequencies larger than 500 Hz, the acceleration of the diaphragm will exceed g.

3. The scale of a spring balance that reads from 0 to 15.0 kg is 12.0 cm long.
A package suspended from the balance is found to oscillate vertically with a
frequency of 2.00 Hz. (a) What is the spring constant? (b) How much does the
package weigh? [HRW5 16-6]

(a) The data in the problem tells us us that the spring within the balance increases in length
by 12.0 cm when a weight of

W = mg = (15.0 kg)(9.80 sm2 ) = 147 N

is pulls downward on its end. So the force constant of the spring must be
F (147 F) N
k= = = 1225 m
x (12 × 10−2 m)

(b) Eq. 4.11 we have the frequency of oscillation of the mass–spring system in terms of the
spring constant and the attached mass. We have the frequency and spring constant and we
can solve to get the mass of the package:
s
1 k k
f= =⇒ m=
2π m 4π 2 f 2
Plug in the numbers:
N
(1225 m )
m= = 7.76 kg
4π (2.00 s )2
2 −1

That’s the mass of the package; its weight is

W = mg = (7.76 kg)(9.80 sm2 ) = 76 N

4. In an electric shaver, the blade moves back and forth over a distance of 2.0 mm
in simple harmonic motion, with frequency 120 Hz. Find (a) the amplitude, (b)
the maximum blade speed, and (c) the magnitude of the maximum acceleration.
[HRW5 16-9]

(a) The problem states that the full distance that the blade travels on each back-and-forth
swing is 2.0 mm, but the full swing distance is twice the amplitude. So the amplitude of the
motion is A = 1.0 mm.
78 CHAPTER 4. OSCILLATORY MOTION

(b) From Eq. 4.6 we have the maximum speed of an oscillating mass in terms of the amplitude
and frequency:

vmax = ωA = 2πfA = 2π(120 s−1 )(1.0 × 10−3 m) = 0.75 ms

(c) From Eq. 4.6 we also have magnitude of the maximum acceleration of an oscillating mass
in terms of the amplitude and frequency:

amax = ω 2 A = (2πf)2 A = 4π 2 (120 s−1 )2 (1.0 × 10−3 m) = 570 sm2

5. The end of one of the prongs of a tuning fork that executes simple harmonic
motion of frequency 1000 Hz has an amplitude of 0.40 mm. Find (a) the maximum
acceleration and (b) the maximum speed of the end of the prong. Find (c)
the acceleration and (d) the speed of the end of the prong when the end has a
displacement of 0.20 mm [HWR5 16-22]

(a) Since we have the amplitude A of the prong’s motion, and we can easily find the angular
frequency ω:
ω = 2πf = 2π(1000 Hz) = 6.28 × 103 s−1
we can use Eq. 4.6 to find the maximum value of a:

amax = ω 2A = (6.28 × 103 s−1 )2 (0.400 × 10−3 m)


= 1.6 × 104 sm2

(b) Likewise, from the same equation we find the maximum speed of the prong’s tip:

vmax = ωA = (6.28 × 103 s−1 )(0.400 × 10−3 m)


= 2.5 ms

(c) Equation 4.7 relates the acceleration a and coordinate x at all times. When the dis-
placement of the prong’s tip is 0.20 mm (half of its maximum) we find

a = −ω 2 x = −(6.28 × 103 s−1 )2(0.20 × 10−3 m) = −7.9 × 103 m


s2

(d) We have already given a relation between |v| (speed) and x in Eq. 4.8. We use it here
to find the speed when x = 0.20 mm:
v
u !2
ux(t)
|v| = ωA 1 −t
A
s
2
0.20 mm

3 −1
= (6.28 × 10 s )(0.40 × 10 −3
m) 1 −
0.40 mm
= 2.2 ms
4.2. WORKED EXAMPLES 79

k m
Frictionless

M
Figure 4.6: Mass M is attached to a spring and oscillates on a frictionless surface. Another block of mass
m is on top!

4.2.2 Mass Attached to a Spring

6. A 7.00 − kg mass is hung from the bottom end of a vertical spring fastened to
an overhead beam. The mass is set into vertical oscillations having a period of
2.60 s. Find the force constant of the spring. [Ser4 13-11]

The formulae in Eq. 4.11 hold even if the mass–spring system oscillates vertically (just
as long as we can neglect the mass of the spring). Then we can solve for the force constant:

m 4π 2 m 4π 2 m
r
T = 2π =⇒ T2 = =⇒ k=
k k T2
and the numbers give us
4π 2(7.00 kg)
k= = 40.9 kg
s2
N
= 40.9 m .
(2.60 s)2
N
The force constant of the spring is 40.9 m .

N
7. Two blocks (m = 1.0 kg and M = 10 kg) and a spring (k = 200 m ) are arranged
on a horizontal, frictionless surface as shown in Fig. 4.6. The coefficient of static
friction between the two blocks is 0.40. What is the maximum possible amplitude
of simple harmonic motion of the spring–block system if no slippage is to occur
between the blocks? [HRW5 16-25]

We first look at what happens when the two blocks oscillate together. In that case it is
legal to regard the mass on the spring as a single mass whose value is M = M + m = 11.0 kg.
We know the spring constant, so using Eq. 4.10 the angular frequency of the motion is
s v
N
200 m
u
k u
ω= = t
= 4.26 s−1
M 11.0 kg
80 CHAPTER 4. OSCILLATORY MOTION

fs fs
m

mg
Figure 4.7: The forces acting on mass m in Example 7. (The force of static friction changes direction and
magnitude during the motion of mass m.)

During the motion, the large mass oscillates with this frequency and so does the small mass
since they move together. But note, the spring is attached only to the large mass; what is
making the small mass move back and forth? The answer is static friction.
We make a diagram of the forces which act on the small mass. This is shown in Fig. 4.7.
We have the force of gravity mg pointing down, the normal force N from the big block
pointing up and also the force of static friction fs , which can point either to the right or
to the left, depending on the current position of m during the oscillation! The magnitude
and direction of the static friction force fs are not constant; the value of fs depends on the
acceleration of the co-moving blocks (assuming there is no slipping so that they are indeed
co-moving).
There is no vertical motion of the small block so clearly

N = mg = (1.00 kg)(9.80 sm2 ) = 9.80 N .

But having the normal force (between the surfaces of the two blocks) we know the maximum
possible magnitude of the static friction force, namely:

fsmax = µs FN = µs mg

and since that is the only sideways force on mass m, from Newton’s 2nd Law, the maximum
possible magnitude of its acceleration — assuming no slipping — is
fsmax µs mg
ano−slip
max = = = µs g .
m m
Now, if the two blocks are moving together and oscillating with amplitude A, then the
maximum value of the acceleration is given by Eq. 4.6, namely amax = ω 2 A, which of course
will get larger if A gets larger. By equating this maximum acceleration of the motion to the
value we just found, we arrive at a condition on the maximum amplitude A such that no
slipping will occur:

ano−slip
max = ω 2 Amax =⇒ µs g = ω 2 Amax
4.2. WORKED EXAMPLES 81

which gives:
µs g (0.40)(9.80 sm2 )
Amax = = = 0.216 m
ω2 (4.26 s−1 )2

4.2.3 Energy and the Simple Harmonic Oscillator

8. A particle executes simple harmonic motion with an amplitude of 3.00 cm. At


what displacement from the midpoint of its motion does its speed equal one half
of its maximum speed? [Ser4 13-23]

The maximum speed occurs in the center of the motion, where there is no potential
energy. So the total energy is given by
2
E = 21 mvmax

At the point(s) where v = 21 vmax the potential energy is not zero; rather it is given by

U = E − 21 mv 2
2
= 12 mvmax − 21 mv 2
vmax 2
 
1 2 1
= 2 mvmax − 2 m
2
1 1
 
2 2
= − mvmax = 83 mvmax
2 8
But we also have from Eq. 4.16 the result
2
mvmax
U(x) = 1
2
x2
A2
And combining these expressions gives the corresponding value of x:
2
1 mvmax 2 2
2 2
x = 38 mvmax
A
Solve for x:
√ √
3 3 3
x = A2
2
=⇒ x=± A=± (3.00 cm) = ±2.60 cm
4 2 2
The mass has half its maximum speed at x = ±2.60 cm.
The problem can also be worked just using Eqs. 4.1 and 4.4. The problem gives no data
about any specific value of t so we are free to choose φ = 0 for simplicity. Then

x(t) = A cos(ωt) and v(t) = −ωA sin(ωt) = −vmax sin(ωt)


82 CHAPTER 4. OSCILLATORY MOTION

and for the times t at which the speed of the mass is half the maximum value, we must have
the condition
sin(ωt) = ± 12 .
But when this is true we have

cos2 (ωt) = 1 − sin2 (ωt) = 1 − 1


4
= 3
4

or √
3
cos(ωt) = ±
2
and that gives √ √
3 3
x = ±A = ±(3.00 cm) = ±2.60 cm
2 2

4.2.4 The Simple Pendulum

9. A simple pendulum has a period of 2.50 s. (a) What is its length? (b) What
would its period be on the Moon, where gMoon = 1.67 sm2 ? [Ser4 13-25]

(a) Using Eq. 4.19 we solve for the length:


s
L L T 2g
T = 2π =⇒ T 2 = 4π 2 =⇒ L=
g g 4π 2

and the numbers give:


(2.50 s)2 (9.80 sm2 )
L= = 1.55 m
4π 2
The length of the pendulum is 1.55 m.
(b) If we take this pendulum to the Moon, its length will be the same, but the acceleration
of gravity will be different. Using the new value of g in Eq. 4.19 we find
s v
L u (1.55 m)
u
TMoon = 2π = 2π t = 6.06 s
gMoon (1.67 sm2 )

The pendulum’s period on the Moon is 6.06 s.

10. If a simple pendulum with length 1.50 m makes 72.0 oscillations in 180 s, what
is the acceleration of gravity at its location? [HRW5 16-59]

We find the frequency f of this pendulum:


72.0
f= = 0.400 Hz
180 s
4.2. WORKED EXAMPLES 83

qmax
L cos qmax
L

2.06 m/s
(a) (b)

Figure 4.8: (a) Pendulum starts with speed 2.06 ms at the bottom of the swing. (b) It attains a maximum
angular displacement θmax .

Then from Eq. 4.19 we can solve for the value of g:


1 g g
r
f= =⇒ (2πf)2 = =⇒ g = 4π 2 f 2 L
2π L L
Plug in the numbers:
g = 4π 2 (0.400 s−1 )2(1.50 m) = 9.47 sm2
The acceleration of gravity at this location is 9.47 sm2 .

11. A simple pendulum having a length of 2.23 m and a mass of 6.74 kg is given
an initial speed of 2.06 ms at its equilibrium position. Assume it undergoes sim-
ple harmonic motion and determine its (a) period, (b) total energy, and (c)
maximum angular displacement. [Ser4 13-59]

The problem is diagrammed in Fig. 4.8 (a).


I will answer the parts of this question in a different order; one reason for this is that
part (c) (maximum value of θ) can clearly be found using energy conservation. Finding the
maximum angular displacement will then give the period.
First off, if we measure height from the bottom of the pendulum’s swing, then in its
initial position it has no potential energy but a kinetic energy equal to

K = 12 mv 2 = 12 (6.74 kg)(2.06 ms )2 = 14.3 J

so the total energy of the system is 14.3 J.


Now when the mass reaches its maximum angular displacement (say, θmax) it is at a
height
ymax = L − L cos θmax = L(1 − cos θmax) .
At that time all of the energy of the particle is potential energy and using energy conserva-
tion, we can solve for θmax :

E = mgymax = mgL(1 − cos θmax) = 14.3 J


84 CHAPTER 4. OSCILLATORY MOTION

(14.3 J) (14.3 J)
(1 − cos θmax) = = = 9.71 × 10−2
mgL (6.74 kg(9.80 sm2 )(2.23 m)
cos θmax = 1 − 9.71 × 10−2 = 9.03 × 10−1
θmax = 25.4◦ = 0.444 rad
This is the exact answer for θmax . Now, one might wonder if 25.4◦ is small enough so that
our calculation of the period of the motion is very accurate, but we forge on anyway!
Now, we are given the linear speed at the bottom of the swing, but the pendulum’s
(harmonic) motion has to do with its angle. We need to relate the two.
From Eq. 1.10 we can get the angular velocity of the mass at the bottom of the swing:
(2.06 ms )
!
dθ v0
= = = 0.924 rad
s
dt 0
L (2.23 m)

But from 4.20 we have θ(t) = θmax cos(ωt + φ) so that the angular velocity of the pendulum
at all times is

= −ωθmax cos(ωt + φ)
dt
so that the maximum angular speed (namely at the bottom of the swing) is
!

= ωθmax
dt max

and this is the same as the 0.924 rad


s
found above. So we can get ω:

(0.924 rad
!
dθ rad )
= 0.924 s
= ωθmax =⇒ ω= s
= 2.08 rad
s
dt max
(0.444 rad)

(It is true that the units don’t look right on that last one, but keep in mind that “radian”
is really dimensionless.)
Having ω, the angular frequency of the pendulum’s oscillations, we go on to get the
period:

T = = 3.02 s
ω
Summing up what problem asked for, we have:

(a) T = 3.02 s (b) E = 14.3 J (c) θmax = 0.444 rad

4.2.5 Physical Pendulums

12. A physical pendulum in the form of a planar body moves in simple harmonic
motion with a frequency of 0.450 Hz. If the pendulum has a mass of 2.20 kg and
the pivot is located 0.350 m from the center of mass, determine the moment of
inertia of the pendulum. [Ser4 13-33]
4.2. WORKED EXAMPLES 85

From Eq. 4.22 we have an expression for the frequency of a “physical pendulum”:
s
1 1 Mgd
f= =
T 2π I
We have all the values that we need to solve for I:
Mgd Mgd
f2 = =⇒ I=
4π 2I 4π 2 f 2
Plug in the numbers:

(2.20 kg)(9.80 sm2 )(0.350 m)


I= = 0.944 kg · m2
4π 2(0.450 s−1 )2

13. A thin disk of mass 5 kg and radius 20 cm is suspended by a horizontal axis


perpendicular to the disk through the rim. The disk is displaced slightly from
equilibrium and released. Find the period of the subsequent simple harmonic
motion. [Tip4 14-57]

We need to find the moment of inertia of the disk when it rotates around an axis at the
rim of the disk. For this, we use the Parallel Axis Theorem of Chapter 1. With ICM given
by ICM = 21 MR2 and recognizing that in this problem, the axis has been shifted a distance
D = R away from the center of mass, we find:

I = ICM + MD2 = 12 MR2 + MR2 = 23 MR2

Then we can find the period of the oscillatory motion from Eq. 4.22. Note, the distance of
the axis from the center of mass (called D in that formula) is R:
v v
s u3 s
I u MR2 3R u 3(0.20 m)
u
t2
T = 2π = 2π = 2π = 2π t = 1.1 s
Mgd MgR 2g 2(9.80 sm2 )

The period of this pendulum is 1.1 s. Interestingly enough, the answer did not depend on
the mass of the disk.
86 CHAPTER 4. OSCILLATORY MOTION

You might also like