0% found this document useful (0 votes)
23 views

Physics Full Lecture Notes (2021)

Uploaded by

aaronlungutsg
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
23 views

Physics Full Lecture Notes (2021)

Uploaded by

aaronlungutsg
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 193

The University of Zambia

School of Natural sciences


Department of Physics
PHY 1010
Lecture 1
Physical Quantities and Units in Physics
Mr. Gift L. Sichone
Phone : +260764036560
Email : [email protected]
May 11, 2020

Introduction
The measurement and recording of quantities is central to the whole of
Physics. The skill of making a reasonable estimate of a physical quantity
is very useful for any Physicist.
This lecture introduces the SI System of Units, which provides a universal
framework of measurement that is common to all scientists internationally.

Learning Outcomes
The student should be able to:
1. understand the nature of all physical quantities i.e. any physical
quantity consists of a numerical magnitude (or size) and a unit;
2. make reasonable estimates of physical quantities included within
the syllabus;
3. recall SI base quantities and their units;
4. express derived units as products or quotients of the SI base units and
use their named units as listed in this syllabus;
5. use the power prefixes and their symbols to indicate decimal sub-multiples
or multiples of both SI base and derived units;

1
Table 1: The SI base quantities and their units

SI Base Quantity SI Base Quantity Symbol Unit Name Unit Symbol


mass m kilogram kg
length l metre m
time t second s
electric current I ampere A
temperature T kelvin K
amount of substance mole mol
light intensity I candela cd

6. make and record measurements of physical quantities in the laboratory;


7. understand Avogadro’s constant and use molar masses.
8. distinguish between scalar and vector quantities and give examples of each;
9. represent a vector as two perpendicular components;
10. add and subtract coplanar vectors.

SI Base Quantities and Units


By convention, modern scientists and engineers use the International System
of Units ( SI - for Systeme Internationale d’Unites). The SI System of Units
uses 7 base quantities and various derived quantities obtained from these
base quantities. The base quantities used in the SI System of Units are mass,
length, time, electric current, temperature, amount of substance and
light intensity. These SI base quantities and their units are listed in Table 1.

Definitions of SI Base Quantities and Units prior to May


2019
Mass
mass refers to an amount of matter with no definite shape. The SI base unit
for mass is kilogram abbreviated kg. 1 kilogram is equal to the mass
of the international standard kilogram, officially known as the International
Prototype Kilogram (IPK). The IPK shown in Figure 1 is a mixture of
platinum and iridium and is held at the International Bureau of Weights and
Measures (website: https://www.bipm.org) in Paris, France. All other masses
were defined by comparing with this metal cylinder.

Time
time refers to a measured period during which an action, process or condition
exits or continues. The SI base unit for time is second. 1 second is the duration

2
Figure 1: The International Prototype Kilogram (IPK) and Standard metre
held at BIPM in Paris, France. The standard metre is made to be exactly the
length that light could travel in 1/299 792 458 of a second.

of 9 192 631 770 periods of the radiation corresponding to the transition between
two hypefine levels of ground state of the caesium-133 atom (see Figure 2).

Length
length refers to the extent of something from end to end. The SI base unit
for length is the metre abbreviated m. 1 metre is defined as the length light
1
travels in vacuum during a time interval of of a second. Figure 1
299792458
shows an example of a standard metre bar held at the International Bureau of
Weights and Measures in Paris.

Electric current
electric current (see Figure 3 top panel) refers to the rate of flow of elec-
tric charge past a point or region. The SI base unit for electric current is
Ampere abbreviated A. 1 Ampere is defined as a constant electric current
which, if maintained in two straight parallel conductors of infinite length, of
negligible cross-section, and placed 1 m apart in vacuum, would produce be-
tween the conductors a force equal to 2 × 10−7 newton per metre of length (see
Figure 3 bottom panel).

Temperature
temperature refers to the degree of hotness or coldness measured on a definite
scale. The SI base unit for temperature is kelvin abbreviated K. 1 kelvin is
1
the fraction of of the thermodynamic temperature of the triple point
273.16

3
Figure 2: The caesium-133 atom is used to define a second. When the outermost
electron in caesium transitions (or moves) between any two hyperfine electron
energy levels, it emits electromagnetic radiation.

4
Figure 3: top panel : An electric current jumping from one conductor to another
conductor through an air gap. bottom panel : Definition of an Ampere.

5
of water. The triple point is defined as the temperature and pressure at which
a substance can exist in equilibrium in the liquid, solid and gaseous states. The
triple point of water is at 0.01◦ C (or 273.16 K) and 4.58 mm (or 611.2 Pa) of
mercury. The triple point is used to calibrate thermometers.

Figure 4: Water at the triple point can co-exist as a solid, liquid and gas. The
triple point of water is used to define the kelvin, the SI base unit for temperature.

Amount of substance
The SI base unit for amount of substance is the mole abbreviated mol. 1
mole is defined as the amount of substance which contains as many elementary
particles (e.g., atoms, molecules, ions, or electrons) as they are atoms in 0.012 kg
(or 12 grams) of carbon-12.

Light Intensity
light intensity refers to the amount of visible light that is emitted in 1 second
(also known as Luminosity, denoted L) per solid angle,Ω. The SI units for
Luminosity is Watt abbreviated W.
The solid angle Ω is defined as the ratio of the area A substended by a cone
to the square of distance r of the source of light. The SI unit for solid angle is
steradians abbreviated sr The SI base unit for light intensity is the standard
candle or candela abbreviated cd (see Figure 6). 1 candela is defined as 1
Watt per steradian.

6
Figure 5: 1 mole of different substances has different masses but has the same
number of elementary particles i.e. 6.02214086 × 1023 particles per mol. This
number is denoted NA is called Avogadro’s number.

Figure 6: The definition of a standard candle or candela, the SI base unit for
light intensity.

7
Table 2: Some well known SI derived quantities and their units

SI Derived Quantity SI Derived Quantity Symbol Unit Name Unit Symbol Base Units Equivalent
area A m2
volume V m3
density ρ kg m−3
speed v m s−1
momentum p kg m s−1
acceleration a m s−2
force F newton N kg m s−2
pressure P pascal Pa kg m−1 s−2
energy (or work) E or W joule J kg m2 s−2
power P watt W kg m2 s−3
frequency ν hertz Hz s−1
charge Q coulomb C As
voltage V volt V kg m2 s−3 A−1
resistance R ohm Ω kg m2 s−3 A−2

SI Derived Quantities and their Units


The SI derived quantities enable us to measure more than the basic quantities of
length, time, mass and etc. Table 2 gives some of the most common SI derived
quantities, their units and equivalent SI base units.

Definitions of SI derived quantities and units


Area

area = length · length


A = l2

Volume

volume = length · length · length


V = l3

Density

mass
density =
volume
M
ρ=
V

8
Speed

distance covered
speed =
time taken
d
v=
t

Linear Momentum

momentum = mass · speed


p = mv

Acceleration

final speed − initial speed


acceleration =
time
v−u
a=
t

Force

force = mass · acceleration


F = ma

Pressure

force
pressure =
area
F
P =
A

Work or Energy

work = force · distance


W = Fd

Power

work
power =
time
W
P =
t

9
Frequency

1
frequency =
Period
1
ν=
T
Period is the time to complete one oscilation or vibration.

Electric Charge

charge = electric current · time


Q = It

Voltage

power
voltage =
electric current
P
V =
I

Resistance

power
resistance =
electric current · electric current
P
R= 2
I

Standard Power Prefixes


Sometimes the values we have to work with for some quantities mean that
the numbers involved are extremely large or small. To address this dilemma,
scientists have made an easier system for writing very large and small numbers
by adding a prefix to the unit which tells us that it has been multiplied by a
very large or small amount. Table gives the prefixes used with SI units.

The Avogadro Constant, NA


The Avogadro Constant, usually denoted NA is the number of atoms in
0.012 kg of carbon-12. The Avogadro constant has a value NA = 6.02214086 ×
1023 atoms mol−1 . Multipying NA by the number of moles of a substance gives
you the number of atoms in sample.
For example, 1 mole of any substance has 6.02214086 × 1023 particles. On a
more informative note,

10
Table 3: Prefixes with SI units

Factor Name Symbol Factor Name Symbol


101 deca da 10−1 deci d
102 hecto h 10−2 cent c
103 kilo k 10−3 mili m
106 mega M 10−6 micro µ
109 giga G 10−9 nano n
1012 tera T 10−12 pico p
1015 peta P 10−15 femto f
1018 exa E 10−18 atto a
1021 zetta Z 10−21 zepto z
1024 yotta Y 10−21 yocto y

6.02214086 × 1023 atoms


0.5 mol = 0.5 mol×NA = 0.5 mol× = 3.01107043×
mol
1023 atoms
6.02214086 × 1023 atoms
2 mol = 2 mol × NA = 2 mol × = 1.204428172 ×
mol
1024 atoms
6.02214086 × 1023 atoms
5 mol = 2 mol × NA = 5 mol × = 3.01107043 ×
mol
1024 atoms

Scalars and Vectors


SI base and derived quantities may be divided into two groups:
scalars - these are physical quantities that have magnitude (size) only.
Examples are mass, length, time, energy, temperature and speed.

vectors - these are physical quantities that with both magnitude and
direction. Examples are force, velocity, acceleration and momentum.
Scalars may be added or subtracted together by simple arithmetic. When
vectors are added or subtracted, the direction of the vector must be considered.
A vector may be represented by a line, the length of the line being the
magnitude of the vector. The direction of the arrow indicates the direction of
the vector.

Addition of vectors
The sum of adding two vectors is referred to as the resultant of the two vectors.

11
Vectors acting in the same line
The resultant of two vectors acting along the same line is the sum of the
magnitudes of the two vectors. The direction of the resultant vector remains
unchanged. If one of the two vectors acting along the same line is acting in the
opposite direction, then its magnitude is given a negative sign.

Figure 7: top panel : Adding two vectors acting along the same line in the same
direction. bottom panel : Adding two vectors acting along the same line but in
different direction.

Vectors acting in different directions


If two vectors are acting in different differentions, a traingle of vectors is used
to find the resultant vector.

Components of vectors
It is often necessary to find the components of a vector, usually in two di-
rections at right angles to each other. This process is called resolution of a
vector. The directions are generally chosen to be the x and y-axes.
The component of a vector along any direction is the magnitude of the vector
multiplied by the cosine of the angle between its direction and the direction of
the vector. A component is the effective value of a vector along a particular
direction.
Usually, we resolve the components of a vector along the x and y-axis di-
rection. The x-component and y-component of a vector A denoted Ax and Ay
respectively are given by

12
Figure 8: Adding two vectors acting along different directions with the aid of a
triangle of vectors

Table 4: Resolving a vector into its components

vector x -component y-component


A Ax = A cos θ Ay = A cos(90 − θ) = A sin θ

The magnitude of A denoted |A| is given by


q
|A| = A2x + A2y

The direction of vector A denoted by θ is given by


Ay
tan θ =
Ax

Consider vectors A, B and C which have magnitudes 20 m, 30 m and 10 m


respectively. The vector A makes 45◦ angle anticlockwise with the positive x−
axis. The vector B makes a 120◦ angle anticlockwise with the positive x− axis.
The vector C lies along the negative x−axis.
These vectors A, B and C can be added or subtracted as shown in Table 5.

13
Table 5: Addition of vectors acting in different directions

vector x -component y-component


A Ax Ay
B Bx By
C Cx Cy
A+B Ax + B x Ay + By
A−B Ax − B x Ay − By
B−A B x − Ax B y − Ay
A+B+C Ax + Bx + Cx Ay + By + Cy
A+B−C Ax + Bx − Cx Ay + By − Cy

Table 6: Addition of vectors acting in different directions

vector x -component y-component


A 20 cos 45◦ = 14.14 m (20) sin 45◦ = 14.14 m
B 30 cos 120◦ = −15 m 30 sin 120◦ = 25.98 m
C 10 cos 180◦ = −10 m 10 sin 180◦ = 0
A+B −0.86 40.12
A−B 29.14 −11.84
B−A −29.14 −29.14
A+B+C −10.86 40.12
A+B−C 9.14 40.12

14
The University of Zambia
School of Natural Sciences
Department of Physics
PHY 1010
Lecture 2
Scalars and Vectors
Mr. Gift L. Sichone
February 28, 2021

1 Introduction
This lecture will introduce two categroies in which all physical quantities can be
categorised called scalars and vectors. Scalars are those physical quantities
that only have magnitude while vectors have both magnitude and direction.
The lecture will focus on decomposition of a vector into two perpendicular scalar
quantities called components and how to add and subtract vectors.

2 Learning Outcomes
By the end of this lecture, the student should be able to:

1. define a scalar quantity and give examples of scalars;


2. define a vector quantity and give examples of vectors;
3. decompose or resolve a vector quantity into two scalar perpendicular com-
ponents;

4. add and subtract coplanar vectors.

3 Scalars and Vectors


SI base and derived quantities may be divided into two groups: scalars and vec-
tors. Scalars are physical quantities that have magnitude or size only. Vectors

1
are those physical quantities that have magnitude and direction. Mathemati-
~ or A. A vector may also be represented
cally, a vector quantity A is denoted as A
pictorially as shown in Figure 1 by a line with an arrow at the end of the line.
The length of the line represents the magnitude of the vector while the arrow
indicates the direction of the vector. Table 1 shows some examples of scalars
and vectors.

Figure 1: A pictorial representation of a vector

Scalars may be added or subtracted together by simple arithmetic as follows:

5 kg + 3 kg = 8 kg

25 s − 13 s = 12 s

25 m + 50 m = 75 m

1.5 mol + 0.7 mol = 2.2 mol

4 Vectors Addition
When vectors are added or subtracted, the direction of the vector must be
considered. The sum of adding vectors is referred to as the resultant. We
only consider vector addition because vector subtraction is a form of vector
addition.

4.1 Adding vectors acting in the same line


The resultant of adding two vectors acting along the same line is the sum of
the magnitudes of the two vectors. The direction of the resultant vector remains
unchanged. If one of the two vectors acting along the same line is acting in the
opposite direction, then its magnitude is given a negative sign.

2
Table 1: List of scalars and vectors

scalars vectors

mass displacement
length velocity
time acceleration
electric current force
temperature pressure
amount of substance momentum
light intensity torque
area electric field
volume magnetic field
density
speed
energy
work
power
frequency
charge
voltage
resistance

4.2 Adding vectors acting along different lines of action


If two vectors are acting in different directions are added, the resultant vector
is found using a traingle of vectors as shown in Figure 3.

5 Components of Vectors
It is often necessary to find the components of a vector, usually in two
directions at right angles to each other. This process is called resolving or
decomposing a vector. The directions along which the components are found
is generally chosen to be the x and y-axes as shown in Figure 4.
The component of a vector along any direction is the magnitude of the vector
multiplied by the cosine of the angle between its direction and the direction of
the vector. A component is the effective value of a vector along a particular
direction. Usually, we resolve the components of a vector along the x and y-axis
direction. The x-component and y-component of a vector A ~ are denoted Ax and
Ay respectively. Table shows the decomposition of vector A ~ into its components
Ax and Ay .

3
Figure 2: Adding two vectors acting along the same line of action

Figure 3: Adding two vectors acting along different directions with the aid of a
triangle of vectors

4
Figure 4: Resolving a vector A into its x and y-components

5
Table 2: Vector resolution or decomposition

vector x -component y-component

~
A Ax = A cos θ Ay = A cos(90 − θ) = A sin θ

~ denoted |A| is given by


The magnitude of A
q
|A| = A2x + A2y

~ denoted by θ is given by
The direction of vector A
Ay
tan θ =
Ax

The angle θ by convention is measured from the positive x-axis in an anti-


clockwise direction.

6 Addition of vectors acting along different di-


rections using table of components
~ B
Consider vectors A, ~ and C ~ which have magnitudes 20 m, 30 m and 10 m
~
respectively. The vector A makes 45◦ angle anticlockwise with the positive x−
~ makes a 120◦ angle anticlockwise with the positive x− axis.
axis. The vector B
~
The vector C lies along the negative x−axis.
~ B
These vectors A, ~ and C ~ can be added or subtracted as shown in Table 3.
~ ~
The magnitude of A + B denoted |A ~ + B|
~ is given by
q
~ + B|
|A ~ = (Ax + Bx )2 + (Ay + By )2

~ + B|
~ =
p
|A (−1 m)2 + (40 m)2


~ + B|
|A ~ = 1601 m2

~ + B|
|A ~ ' 40 m
~+B
The direction of the vector A ~ is given by

6
~,B
Figure 5: The coplanar vectors A ~ and C
~ on the xy-plane.

~+B
Figure 6: The resultant vector A ~

7
Table 3: Addition of vectors acting in different directions

vector x -component y-component

~
A Ax Ay
~
B Bx By
~
C Cx Cy
~+B
A ~ Ax + Bx Ay + B y
~
A−B~ Ax − Bx Ay − B y
~ −A
B ~ B x − Ax B y − Ay
~ ~
A+B+C~ Ax + Bx + Cx Ay + B y + C y
~ ~
A+B−C~ Ax + Bx − Cx Ay + B y − C y

Ay + B y
tan φ =
Ax + B x
40 m
tan φ = = −40
−1 m

φ = tan−1 (−40)

φ ' −89◦

θ = 180◦ + φ = 180◦ + (−89◦ )

θ = 91◦
~−B
The magnitude of A ~ denoted |A
~ − B|
~ is given by
q
~ − B|
|A ~ = (Ax − Bx )2 + (Ay − By )2

~ − B|
~ =
p
|A (29 m)2 + (−12 m)2


~ − B|
|A ~ = 985 m2

~ − B|
|A ~ ' 31.4 m
~−B
The direction of the vector A ~ is given by

8
Table 4: Addition of vectors acting in different directions

vector x -component y-component

~
A Ax = (20 m) cos 45◦ ' 14 m Ay = (20 m) sin 45◦ ' 14 m
~
B Bx = (30 m) cos 120◦ = −15 m By = (30 m) sin 120◦ ' 26 m
~
C Cx = (10 m) cos 180◦ = −10 m Cy = (10 m) sin 180◦ = 0
~+B
A ~ Ax + Bx = −1 m Ay + By = 40 m
~
A−B~ Ax − Bx = 29 m Ay − By = −12 m
~ −A
B ~ Bx − Ax = −29 m By − Ay = 12 m
~ ~
A+B+C~ Ax + Bx + Cx = −11 m Ay + By + Cy = 40 m
~ ~
A+B−C~ Ax + Bx − Cx = 9 m Ay + By − Cy = 40 m

~−B
Figure 7: The resultant vector A ~

Ay − B y
tan φ =
Ax − B x
−12 m
tan φ = = −0.41
29 m

φ = tan−1 (−0.41)

9
φ ' −22.3◦

θ = 360◦ + φ = 360◦ + (−22.3◦ )

θ = 337.7◦

~ −A
Figure 8: The resultant vector B ~

~ −A
The magnitude of B ~ denoted |B
~ − A|
~ is given by
q
~ − A|
|B ~ = (Bx − Ax )2 + (By − Ay )2

~ − A|
~ =
p
|B (−29 m)2 + (12 m)2


~ − A|
|B ~ = 985 m2

~ − A|
|B ~ ' 31.4 m
~ −A
The direction of the vector B ~ is given by

By − Ay
tan φ =
Bx − Ax

10
12 m
tan φ = = −0.41
−29 m

φ = tan−1 (−0.41)

φ ' −22.3◦

θ = 180◦ + φ = 180◦ + (−22.3◦ )

θ = 157.7◦

~+B
Figure 9: The resultant vector A ~ +C
~

~+B
The magnitude of A ~ +C
~ denoted |A
~+B ~ + C|
~ is given by
q
~+B
|A ~ + C|
~ = (Ax + Bx + Cx )2 + (Ay + By + Cy )2

~+B
~ + C|
~ =
p
|A (−11 m)2 + (40 m)2


~+B
|A ~ + C|
~ = 1 721 m2

~+B
|A ~ + C|
~ ' 41.5 m
~+B
The direction of the vector A ~ +C
~ is given by

11
Ay + B y + C y
tan φ =
Ax + Bx + Cx
40 m
tan φ = = −3.64
−11 m

φ = tan−1 (−3.64)

φ ' −74.6◦

θ = 180◦ + φ = 180◦ + (−74.6◦ )

θ = 105.4◦

~+B
Figure 10: The resultant vector A ~ −C
~

~+B
The magnitude of A ~ −C
~ denoted |A
~+B ~ − C|
~ is given by
q
~+B
|A ~ − C|
~ = (Ax + Bx − Cx )2 + (Ay + By − Cy )2

~+B
~ − C|
~ =
p
|A (9 m)2 + (40 m)2


~+B
|A ~ − C|
~ = 1 681 m2

~+B
|A ~ − C|
~ ' 41 m

12
~+B
The direction of the vector A ~ −C
~ is given by

Ay + By − Cy
tan θ =
Ax + B x − C x
40 m
tan θ = = 4.44
9m

θ = tan−1 (4.44)

θ ' 77.3◦

13
The University of Zambia
School of Natural Sciences
Department of Physics
PHY 1010
Lecture 3A
Description of Uniformly Accelerated Motion
Mr. Gift L. Sichone
March 6, 2021

1 Introduction
The movement of objects (living and non-living) is a part of our everyday ex-
perience. So it is important that we are able to study, analyse and predict the
motion of objects in our surroundings. In this course, we are particularly inter-
ested in a type of motion referred to as uniform accelerated motion. This
uniform acelerated motion of objects can easily be studied both graphically and
through mathematical like expressions called equations of motion. The study
of the motion of objects without referring to the cause of the motion is called
kinematics.

2 Learning Outcomes
By the end of this lecture, the student should be able to:

1. define and use distance, displacement, speed, velocity and acceleration;


2. use graphical methods to represent distance, displacement, speed, velocity
and acceleration;
3. determine displacement from the area under a velocity-time graph;
4. determine velocity using the gradient of a displacement-time graph;
5. determine acceleration using the gradient of a velocity-time graph;

1
3 Description of Uniformly Accelerated Motion
3.1 Distance and Displacement
Distance and displacement are closely related in that both physical quantities
are a form of length. Distance denoted d is a scalar while displacement denoted
s is a vector. Displacement refers to the distance moved by an object in a
specific direction, for example, 120 m west. The SI units for distance and
displacement are metres.

3.2 Speed and Velocity


Speed and velocity are also closely related but do not refer to the same physical
quantity. Speed is a scalar while velocity is a vector. Speed only refers to how
fast an object is moving (i.e. rate of movement) and does not say anything
about the direction in which the object is moving. Velocity, on the other hand,
refers to the speed of an object in a specific direction, for example, a car moving
at 150 m/s east.
There are two type of speed or velocity. The first type is called instan-
taneous speed or velocity while the second type is the average speed or
velocity.Instantaneous speed or velocity refers to the speed or velocity of
an object in motion at a specific point in time. The speed limit on the side of
a road shows the maximum speed or velocity a motorist is lawfully allowed to
drive a vehicle in that section of road. Instantaneous speed or velocity is what
traffic police use to determine if a motorist has been overspeeding on a section
of road or not.
Average speed or velocity, on the other hand, is obtained by dividing
the total distance travelled by an object by the total time taken to travel that
distance. Average speed or velocity is the preferred physical quantity we
tend to use to describe how fast an object was moving because it convieniently
gives us only one value of speed to work with. Instantaneous speed or velocity
on the other hand generates different values of speed or velocity at each point
in time which can complicate our ability to analyse and predict the motion of
an object.
The average speed denoted v̄ is defined as follows
total distance covered
average speed =
total time taken

d
v̄ = (1)
t
where

v̄ is the average speed in metres per second (m/s)


d is the total distance covered in metres (m)
t is the total time taken in seconds (s)

2
The average velocity denoted v̄ is defined as follows
total displacement covered
average velocity =
total time taken
s
v̄ = (2)
t
where

v̄ is the average velocity in metres per second (m/s)


s is the total displacement covered in metres (m)
t is the time taken in seconds (s)

Alternatively, average velocity can also be defined in terms of the initial


velocity denoted u and the final velocity denoted v of a moving object as follows:
1
average velocity = (initial velocity + final velocity)
2
1
v̄ = (u + v) (3)
2
where

v̄ is the average velocity in metres per second (m/s)


u is the initial velocity in metres per second (m/s)
v is the initial velocity in metres per second (m/s)

Example 1
An athlete takes 10 s to run a distance of 100 m. Calculate the average speed
of the athlete during this 10 s.

total distance covered


average speed =
total time taken

d
v̄ =
t
100 m
v̄ =
10 s

v̄ = 10 m/s
On average, the athlete covers a distance of 10 m per second. In reality, the
athlete does not actually cover a distance of 10 m every second of the 10 s he
or she covers the 100 m distance.

3
3.3 Acceleration
Acceleration refers to the rate of change of speed or velocity. In this course,
we are interested in a type of acceleration called average acceleration which
is defined as follows:
change in velocity or speed
average acceleration =
time taken
∆v
a= (4)
t
where

a is the average acceleration in metres per second squared (m/s2 )


∆v is the change velocity in metres per second (m/s)
t is the time taken in seconds (s)

The average acceleration can also be expressed in terms of the initial and
final velocity as follows:

final velocity or speed − initial velocity or speed


average acceleration =
time taken

v−u
a= (5)
t
where

a is the average acceleration in metres per second squared (m/s2 )


v is the initial velocity in metres per second (m/s)

u is the initial velocity in metres per second (m/s)


t is the time taken in seconds (s)

Example 2
A car accelerates from rest to 28 m/s in 10 s. Calculate the average acceleration
the car experiences.

initial velocity of car u = 0 m/s


final velocity of car v = 28 m/s

time t = 10 s

4
Table 1: The velocity of an accelerating car in 10 s.

t (s) 0 1 2 3 4 5 6 7 8 9 10
v (m/s) 0 2.8 5.6 8.4 11.2 14 16.8 19.6 22.4 25.2 28

v−u
a=
t
28 m/s − 0 m/s
a=
10 s
28 m/s
a=
10 s

a = 2.8 m/s2
The car experiences an average acceleration of a = 2.8 m/s2 . This means
that the speed or velocity of the car increases by 2.8 m/s every second during
this duration of 10 s. Table 1 shows how the velocity of this accelerating car
changes during these 10 s.

4 Graphical Representation of Uniformly Accel-


erated Motion
The motion of a moving object can usually be displayed graphically. The most
useful motion graph is the velocity-time graph denoted v−t graph. Other motion
graphs of note are distance-time or displacement-time graphs and acceleration-
time graph.
The acceleration of a moving object can be obtained from the v −t graph and
is equal to the slope or gradient of the v −t graph. The distance or displacement
covered by a moving body is equal to the area under the v − t graph.
For example, consider a car that starts from rest and accelerates to 28 m/s
in 10 s. The acceleration of the car is 2.8 m/s2 . Physically, this acceleration
means that velocity of the car increases by 2.8 m/s every second. Table 1 shows
how the velocity of the car changes with time in these 10 s.
From the v − t graph shown in Figure 1, if we pick the points (t3 , v3 ) =
(3 s, 8.4 m/s) and (t8 , v8 ) = (8 s, 22.4 m/s), we can calculate the slope or gradient
as follows:
change in velocity
slope or gradient =
change in time
v8 − v3
slope =
t8 − t3

5
Figure 1: The velocity-time graph of a car accelerating at a = 2.8 m/s2 for 10 s.

22.4 m/s − 8.4 m/s


slope =
8 s−3 s
14 m/s
slope =
5s

slope = 2.8 m/s2


The slope of this v − t graph is equal to the acceleration of the car, that is,
2.8 m/s2 . Every second, the velocity of this car increases by 2.8 m/s.
The total area under the v − t graph in Figure 1 gives the total distance or
displacement covered by the car in 10 s. In this case, the area under the v − t
graph, corresponds to the area of the triangle. The area of a triangle is given
by:
1
area of triangle = × base × height
2
In this case, the base = 10 s and the height = 28 m/s. Therefore, we obtain
an area of triangle given by
1
area of triangle = × 10 s × 28 m/s = 140 m.
2

6
Table 2: The displacement of an accelerating car in 10 s.

t (s) 0 1 2 3 4 5 6 7 8 9 10
s (m) 0 1.4 5.6 12.6 22.4 35 50.4 68.6 89.6 113.4 140

The area under an acceleration-time a − t graph gives the average veloc-


ity. Note that we are only concerned with uniform accelerated motion i.e. the
acceleration is constant and does not change with time.

Figure 2: The s − t graph of a car accelerating at a = 2.8 m/s2 for 10 s.

7
Figure 3: The a − t graph of a car accelerating at a = 2.8 m/s2 for 10 s.

Figure 4: The v − t graph of a car accelerating at a = −2.8 m/s2 for 10 s.

8
Figure 5: The s − t graph of a car accelerating at a = −2.8 m/s2 for 10 s.

Figure 6: The a − t graph of a car accelerating at a = −2.8 m/s2 for 10 s.

9
The University of Zambia
School of Natural Sciences
Department of Physics
PHY 1010
Lecture 3B
Equations of Uniform Accelerated Motion
Mr. Gift L. Sichone
March 6, 2021

1 Introduction
In this lecture, we will derive the equations of motion that are used to describe
uniformly accelerated motion in a straight line.

2 Learning outcomes
By the end of this lecture, the student should be able to:

1. state the five equations of motion that are used to represent uniformly
accelerated motion in a straight line;

2. use the equations of motion used to describe uniformly accelerated motion


to solve problems.

3 Equations of uniformly accelerated motion in


a straight line
The velocity-time (v − t) graph is particularly important for deriving the equa-
tions of uniform accelerated motion because its gradient is equal to the acceler-
ation and the area under the graph is equal to the displacement.

acceleration = slope of velocity time graph

1
Figure 1: The velocity-time graph of a car accelerating at a = 2.8 m/s2 for 10 s.

To obtain the gradient of a v − t graph, we need to define the following


physical quantities:
s - displacement (m)

t - time (s)
u - initial velocity (m/s) at t = 0
v - final velocity (m/s) after t

a - acceleration (m/s2 )

change in velocity
acceleration =
time taken
∆v
a=
t
where

a is the average acceleration in m/s2 ;


∆v is the change in velocity in m/s
t is the time taken in s

2
v−u
a= (1)
t
where

a is the average acceleration in m/s2 ;


v is the final velocity in m/s;
u is the final velocity in m/s;

t is the time taken in s.


This is the first equation of motion used to describe uniform accelerated motion
in a straight line.
Multiplying both side of the equation (1) by t, we obtain:

at = v − u
Moving u to the other side of the equation, we obtain the second equation
of motion used to describe uniform accelerated motion in a straight line.

v = u + at (2)
The area under the v − t graph is equal to the total displacement s distance
d moved. Recall, that we defined average velocity denoted v̄ as follows
1
v̄ = (u + v) (3)
2
Next, we obtain the total displacement s in terms of average velocity v̄ and
time t as follows:

s = v̄t (4)
Substituting for the average velocity v̄, we obtain
1
s= (v + u) t (5)
2
This is the third equation of motion used to describe uniformly accelerated
motion in a straight line.

vt + ut
s=
2
Substituting for v from the first equation of motion, we obtain:

(u + at)t + ut
s=
2

3
ut + at2 + ut
s=
2

2ut + at2
s=
2
1
s = ut + at2 (6)
2
This is the fourth equation of motion used to describe uniformly accelerated
motion in a straight line.
We can re-write equation (1) as:
v−u
t=
a
and substitute it into equation (6), we get.
v−u 1 v−u 2
s = u( ) + a( )
a 2 a
This gives us

uv − u2 1 (v − u)2
s= + a
a 2 a2

uv − u2 (v − u)(v − u)
s= +
a 2a

uv − u2 (v 2 − uv − uv + u2 )
s= +
a 2a

2uv − 2u2 + v 2 − 2uv + u2


s=
2a

v 2 − u2
s=
2a

2as = v 2 − u2

v 2 = u2 + 2as (7)
This is the fifth equation of motion used to describe uniform linearly accelerated
motion in a straight line.

4
4 Uniform accelerated motion due to gravity
Uniform accelerated motion due to gravity is divided into upward motion from
the surface of the Earth and downward motion towards the surface of the Earth.
When an object is launched or thrown upwards, its velocity reduces at a steady
9.8 m/s every second until it reaches its maximum height, where its velocity is
0 m/s for a very short time. Every body thrown vertically upwards therefore
accelerates at -9.8/m s2 .
On the other hand, if an object is dropped from a height, it accelerates
towards the surface of the Earth at a steady 9.8 m/s every second. At the time,
of being droped the object has an initial velocity of 0 m/s and once released will
increase its velocity at a rate of 9.8/m s every second. This means that that the
body will accelerate at 9.8 m/s2 .
In both cases (i.e, upward and downward motion), the five equations of
motion for uniformly accelerated motion still apply because the acceleration is
constant i.e. 9.8 m/s2 . Since our acceleration a = g = 9.8m/s2 , our equations
of motion become:
v−u
g= (8)
t

v = u + gt (9)

1
s= (u + v)t (10)
2
1
s = ut + gt2 (11)
2

v 2 = u2 + 2gs (12)
During the upward motion of a body, the acceleration due to gravity has a
value of g = −9.8 m/s2 . On the other hand, the acceleration due of gravity has
a value of g = 9.8 m/s2 in the downward motion of a body.

Example 3
A bullet from a rifle shot upwards has a muzzle velocity of 550 m/s. Calculate:

(a) the time it takes the bullet to reach its maximum height.
(b) the maximum height reached by the bullet.
(c) the total time of flight of the bullet
(d) the velocity with which the bullet strikes the ground.

5
Solutions
(a) The bullet exits the rifle with a muzzle velocity (i.e. initial velocity) u =
550 m /s. When the bullet reaches its maximum height, its final velocity
v = 0 m/s. Since the bullet is undergoing upward motion, acceleration
due to gravity g becomes -9.8 m/s2 .
We can obtain, the time of flight t from the equation v = u + gt.

v = u + gt

0 m/s = 550 m/s + (−9.8 m/s2 )trise

Dropping off units, we obtain

0 = 550 + (−9.8)trise

0 = 550 − 9.8trise

9.8trise = 550

Dividing both sides by 9.8, we get

550
trise =
9.8

trise = 56.1 s

(b) the maximum height reached by the bullet, smax


Now that we know the rise time of the bullet, trise = 56.1 s and the initial
velocity, u = 550 m/s, we can use the equation of motion
1
s = ut + gt2
2
to get the maximum height denoted smax

1
smax = ut + gt2
2

1
smax = (550 m/s)(56.1 s) + (−9.8 m/s2 )(56.1 s)2
2
Dropping off units, we get

6
1
smax = (550)(56.1) − (9.8)(56.1)2
2

smax ≈ 15 434 m ≈ 15.434 km

The bullet rises to a maximun height of smax ≈ 15.434 km


(c) the total time of flight of the bullet.
At maximum height, the bullet has an initial velocity u = 0 m/s. As the
bullet returns to Earth it gathers velocity as its accelerates at g = 9.8 m/s2 .
The bullet on its downward motion covers a distance of s = 15 434 m.
Therefore, we can work out the time of fall on its downward journey using
1
s = ut + gt2
2
. Substituting the values, we have

1
smax = ut + gt2
2

1
15 434 m = (0 m/s)tfall + (9.8 m/s2 )t2fall
2
Dropping off units, we get
1
15 434 = (0)tfall + (9.8)t2fall
2

15 434 = 0 + 0.5(9.8)t2fall

15 434 = (4.9)t2fall

4.9t2fall = 15 434
Dividing both sides by 4.9, we get

15 434
t2fall = = 3 149.8
4.9
Getting the square root on both sides, we obtain

tfall = 3 149.8

tfall = 56.1 s

7
the fall time on the downward motion is again tfall = 56.1 s.
Therefore, the total time of flight for both the upward and downward
motion of the bullet is

tflight = trise + tfall

tf = 56.1 s + 56.1 s

tf = 112.2 s

(d) the velocity with which the bullet strikes the ground.
Now that we know the fall time of the bullet on its downward motion,
tfall = 56.1 s, and its initial velocity at maximum height u = 0 m/s, we
can calculate the final velocity with which the bullet strikes the ground,
v.

v = u + gt
v = 0 m/s + 9.8 m/s2 × 56.1 s

Dropping off units, we get

v = 0 + 9.8(56.1)

v = 549.78

v ∼ 550 m/s
The bullet strikes the ground with a velocity of 550 m/s. This is the same
as the muzzle velocity with which the bullet is fired upwards from the gun.

8
The University of Zambia
School of Natural Sciences
Department of Physics
PHY 1010
Lecture 3C
Projectile Motion
Mr. Gift L. Sichone
March 6, 2021

1 Introduction
This lecture introduces the motion of projectiles under the influence of gravity.
A projectile is a body that has been projected by an external force and continues
in motion by its own inertia. In this lecture, we will learn how to describe the
motion of a projectiles which is characterized to movement in two perpendicular
directions.

2 Learning outcomes
By the end of this lecture, the student should be able to:

1. describe in words the motion of a projectile due to a uniform velocity in


horizontal direction and a uniform acceleration in the vertical direction;
2. calculate the rise time of a projectile to maximum height;
3. calculate the maximum height the projectile rises;

4. calculate the fall time of a projectile from maximum height;


5. calculate the flight time of a projectile i.e. sum of rise time and fall time;
6. calculate the horizontal distance travelled by a projectile i.e. the range

7. calculate the velocity with which a projectile strikes the ground.

1
Table 1: Decomposition of projectile muzzle velocity u = 550 m/s at θ = 45◦
into ux and uy components

vector x -component y-component

~u ux = (550 m/s) cos 45◦ = 389 m/s uy = (550 m/s) sin 45◦ = 389 m/s

3 Projectile Motion of a shell


Consider a shell fired from an anti-aircraft gun with a muzzle velocity of 550 m/s
at an angle of 45◦ with the ground. The muzzle velocity of the shell is a vector
as it has magnitude of 550 m/s and direction of 45◦ from the horizontal. In the
case of projectiles, we are generally interested in finding out how high the shell
will go up into the sky i.e. maximum height and how far the shell will travel
horizontally i.e. the range.
Projectiles always carries out two types of motion at the same time. The first
type of motion is vertical (i.e. up and down) motion along the y-direction and
the second type of motion is horizontal motion along the x-direction. These
two types of motions are connected to one another by the flight time of the
projectile. The flight time is defined as the time it takes the projectile to rise
from the ground, reach maximum height and then return to the ground. This
is the same time it takes the projectile of travel the horizontal distance referred
to as the range.
Another important point to note is that for a projectile to reach maximum
height, it upward or vertical velocity at maximum height must be equal to
0 m/s. Otherwise, the projectile will continue to rise. Also at maximum height,
the projectile is not stationary but continues to move along the horizontal with
a non-zero velocity.
To solve any problem of projectile motion, we first need to decompose the
initial muzzle velocity u into its perpendicular components. We first need to
find out the magnitude of the initial muzzle velocity along the the x-direction
(i.e. x-component) denoted ux and y-direction ( i.e. y-component) denoted uy .
Table 1 shows the decompostion of the initial muzzle velocity u of the shell fired
from an anti-aircraft gun into ux and uy .
The second step is to consider only the vertical motion of the projectile
along the y−direction and ignore the horizontal motion of the projectile. In the
y-direction, the velocity of the projectile will continue to decrease at a steady
9.8 m/s every second i.e. 9.8 m/s2 until it becomes 0 m/s at maximum height,
smax . Concerning this upward motion of the projectile, we can find out the
maximum height as well as the time it will take the projectile to reach maximum
height using our equations of motion in a straight line.

2
Figure 1: Projectile motion of a shell as a function of flight time

3.1 Calculating the rise time of a projectile


We know that to reach maximum height denoted smax , the initial velocity in
the y-direction i.e. uy = 389 m/s has to decrease to a final velocity uy of 0 m/s
at a steady rate of 9.8 m/s every second i.e., g = -9.8 m/s2 . Therefore, we can
use the equation

v = u + gt
To obtain the time to rise to maximum height, rise time, denoted, trise . We
get

vy = uy + gtrise
where

vy = 0 m/s;
uy = 389 m/s
g = −9.8 m/s2
trise is the rise time.

Substituting these values in the above equation, we obtain

0 m/s = 389 m/s + (−9.8 m/s2 )trise


Dropping the units, we get

3
0 = 389 + (−9.8)trise

0 = 389 − 9.8trise

9.8trise = 389
Dividing both sides by 9.8, we get
389
trise =
9.8

trise = 39.7 s
Therefore, it takes the projectile trise = 39.7 s to reach the maximum height.

Figure 2: Projectile motion of a shell as a function of range.

3.2 Calculating maximum height reached by projectile


From the calculated rise time to reach maximum height trise = 39.7 s, we can
calculate the maximum height, smax , reached using the following equation
1
s = ut + gt2
2

4
We know uy = 389 m/s, g = −9.8 m/s2 and trise = 39.7 s.
1
smax = uy trise + gt2rise
2
Substituting, we obtain
1
smax = (389 m/s) × (39.7 s) + × (−9.8 m/s2 ) × (39.7 s)2
2
Dropping off the units, we get

smax = 389 × 39.7 − 0.5 × 9.8 × (39.7)2

We get the maximum height , smax as

smax = 7 720 m.

The projectile rises to a maximum height of 7 720 m.

3.3 Calculating time of fall from maximum height


After obtaining the maximum height, we can proceed to calculate fall from
maximum height, tfall . We know that at maximum height, smax , in the y-
direction, the velocity of the projectile is 0 m/s and that the projectile will have
to fall through a distance of 7 720 m to strike the ground. During this fall, the
velocity of the projectile increases by 9.8 m/s every second. i.e. g = 9.8 m/s2 .
We can therefore, get the time of fall, tfall from the equation
1
s = ut + gt2
2
In the y-direction, this equation becomes
1
smax = uy tfall + g t2fall
2
We know that uy = 0 m/s, g = 9.8 m/s2 and smax = 7 720 m. We are
looking for tfall . Substituting, we obtain
1
7 720 m = (0 m/s) × tfall + × (9.8 m/s2 ) × t2fall
2
Dropping off the units, we get
1
7 720 = 0 × tfall + × (9.8) × t2fall
2

7 720 = 0 + 4.9 t2fall

7 720 = 4.9 t2fall

5
4.9 t2fall = 7 720
Dividing both sides of the equation by 4.9, we obtain
7 720
t2fall = = 1 575.5
4.9
Square rooting both sides of the equation, we obtain
q √
t2fall = 1 575.5

tfall = 39.7 s
The fall time ,tfall , is 39.7 s. as can be seen in this case, the time it takes
the projectile to rise to its maximum height and the time it takes to fall are the
same.

trise = tfall = 39.7 s

Figure 3: The projectile velocity of a shell as a function of flight time.

6
3.4 Calculating flight time of a projectile
We can obtain the flight time denoted tf as a sum of the time to rise to maximum
height,trise , and the time of fall,tfall .

tf = trise + tfall

tf = 39.7 s + 39.7 s

tf = 79.4 s

3.5 Calculating vertical velocity with which projectile strikes


the ground
We can also calculate the vertical final velocity vy with which the projectile
strikes the ground. This can be done using the time of fall, tf all =39.7 s and the
initial velocity of the projectile at maximum height, uy = 0 m/s. We can obtain
vy from the equation

v = u + gt
In the y-direction, we get

vy = uy + gt2
Substituting for uy = 0 m/s and t2 = 39.7 s, we get

vy = (0 m/s) + (9.8 m/s2 ) × (39.7 s)

Dropping the units we get,

vy = 0 + 9.8 × 39.7

vy = 389 m/s
From this value of vy = 389 m/s, we see that the projectile strikes the ground
with the same velocity with which it rose up with.

uy = vy = 389 m/s
We have found out all the key parameters about the motion of the projectile
in the y-direction.

7
3.6 Calculating the range of the projectile
We can now consider motion in the x-direction. You need to be aware that
there is no acceleration i.e. ax =0 m/s in the x-direction. Therefore, the initial
velocity of the projectile in the x-direction is the same as the final velocity of
the projectile in the x-direction, i.e., ux = vx = 389 m/s. The time of flight
of the projectile has been found to be 79.4 s. Therefore, we can calculate the
range as

range = ux × tf

range = 389 m/s × 79.4 s

range = 30 886.6 m

range ∼ 30.9 km

Figure 4: The projectile velocity of a shell as a function of flight time.

3.7 Calculating velocity with which projectile strikes the


ground
When the projectile strikes the ground after flying for 79.4 s, it has a final
velocity with vx = 389 m/s and vy = 389 m/s. The magnitude of the velocity

8
with which the projectile strikes the ground is given by
q
v = vx2 + vy2

.
Therefore, we get a magnitude of of
p
v = (389 m/s)2 + (389 m/s)2

v = 550 m/s
This projectile strike velocity is the same as the projectile muzzle velocity from
the anti-aircraft gun. So we conclude that the strike velocity and muzzle velocity
in this case are the same.

9
The University of Zambia
School of Natural Sciences
Department of Physics
PHY 1010
Lecture 4A
Newton’s Laws of Motion
Mr. Gift L. Sichone
April 27, 2021

1 Introduction
The motion of any object is governed by forces that act on the object. This
lecture introduces Newton’s laws of motion, which are fundamental to under-
standing the connection between forces and motion.

2 Learning Outcomes
By the end of this lecture, the student should be able to:

1. understand that mass is a property of a body that resists change in motion;

2. state and apply each of Newtons’s Laws of Motion;


3. recall the relationship F = ma and solve problems using it, appreciating
that the acceleration a and the resultant force F always are in the same
direction;

4. describe and use the concept of weight as the effect of a gravitational field
on a mass and recall that the weight of a body is equal to the product of
its mass and the acceleration of free fall, FW = mg;

1
3 Newton’s Laws of Motion
Mass
The mass of an object is a measure of the inertia of the object. Inertia is the
tendency of a body at rest to remain at rest, and of a body in motion to continue
moving with unchanged velocity. The kilogram (kg) is the SI unit of mass. 1 kg
is the mass of an object whose mass is equal to the International Prototype
Kilogram or standard kilogram held at the International Bureau of Weights
and Measures in Paris, France. The masses of all other objects are compared
to the standard kilogram.

Force
In mechanics, force is that which changes the velocity of an object. Force
is a vector quantity, having magnitude and direction. An external force is
one whose source lies outside of the system (or body) being considered. The
net external force acting on an object causes the object to accelerate in the
direction of that force. The acceleration is proportional to the force and inversely
proportional to the mass of the object. The newton (N) is the SI unit of force.
1 N is that resultant force which will give a 1 kg mass an acceleration of 1 m/s2
.

Newton’s First Law of motion


An object at rest will remain at rest; an object in motion will continue in motion
with constant velocity until it is acted upon by an external force. Force is the
changer of motion.

Newton’s Second Law of Motion


If the resultant (or net), force F acting on an object of mass m is not zero, the
object accelerates in the direction of the force. The acceleration a is proportional
to the force F and inversely proportional to the mass of the object m.
F
a= , F = ma (1)
m
The acceleration a has the same direction as the resultant force F .
The vector equation F = ma can be written in terms of components as
X X
Fx = max , Fy = may (2)
P P
where the Fx and Fy are the resultant forces are the components of the
external forces acting on the object in the x and y-directions. The resulting
accelerations in the x and y-directions are given by ax and ay .
q
a = a2x + a2y (3)

2
Newton’s Third Law of Motion
For each force exerted on one body, there is an equal, but oppositely directed,
force on some other body interacting with it. This is often called the Law of
Action and Reaction. Notice that the action and reaction forces act on the
two different interacting objects.

4 The Law of Universal Gravitation


When two masses m1 and m2 gravitationally interact, they attract each other
with forces of equal magnitude. For point masses (or spherically symmetric
bodies), the attractive force FG is given by
m1 m2
FG = G (4)
r2
where r is the distance between mass centers, and where G = 6.67 ×
10−11 N m2 /kg, where FG is in newtons, m1 and m2 are in kilograms, and
r is in meters.
To calculate the force of gravity, experienced here on Earth, we let m1 be
the mass of the Earth and denote it as mE = 5.97219 × 1024 kg. We also let m2
be the masss of an object on the surface of the Earth and denote it as just m.
In this cases, the distance between the centre of the Earth and the object on the
surface of the Earth is just the radius of the Earth denoted rE = 6 371 000 m.
Re-writing the Law of Universal Gravitation, we get
mmE
FG = G 2 (5)
rE
 
mE
FG = m G 2 (6)
rE
 
mE
The term G 2 is a constant on Earth and is denoted as g
rE
mE
g=G 2 (7)
rE

5.97219 × 1024 kg
 
g = (6.67 × 10−11 N m2 /kg2 )
(6 371 000 m)2

g ' 9.81 m/s2 (8)


The g is our acceleration due to gravity here on Earth.
Therefore, the Law of Universal Gravitation can be expressed as

FG = mg (9)
Gravity on Earth causes objects to accelerate downwards towards the center
of Earth by 9.8 m/s2 .

3
5 Types of forces
5.1 Weight
The weight of an object is the gravitational force acting downward on the object.
On the Earth, weight is the gravitational force exerted on the object by the
planet. Its SI units of weight are newtons
An object of mass m falling freely toward the Earth is subject to only one
force, the pull of gravity, which we call the weight FW of the object. The object’s
acceleration due to FW is the free-fall acceleration g. Therefore, F = ma
provides us with the relation between F = FW , a = g, and m; it is FW = mg.
Because, on average, g = 9.8 m s−2 on Earth, a 1.0 kg object weighs 9.8 N at
the Earth’s surface.

5.2 Normal force


The Normal force FN on an object that is being supported by a surface is the
component of the supporting force that is perpendicular to the surface.

Figure 1: Forces acting on a box on a horizontal surface.

5.3 Tensile Force


The tensile force acting on a string or chain or tendon is the applied force tending
to stretch it. The magnitude of the tensile force is the tension denoted FT .

4
Figure 2: Forces acting on a box supported by a rope

5.4 Friction Force


The friction force Ff is a tangential force acting on an object that opposes the
sliding of that object on an adjacent surface with which it is in contact. The
friction force is parallel to the surface and opposite to the direction of motion
or of impending motion. Only when the applied force exceeds the maximum
static friction force will an object begin to slide.

6 Coefficients of Friction
6.1 Coefficient of static friction
The coefficient of static friction µs is defined for the case in which one surface
is just on the verge of sliding across another surface. It is
static friction force
coefficient of static friction =
normal force
Ff (max)
µs = (10)
FN

static friction force = coefficient of static friction × normal force

Ff (max) = µs · FN (11)
where the maximum friction force occurs when the object is just on the verge
of slipping but is nonetheless at rest.

5
6.2 Coefficient of kinetic friction
The Coefficient of kinetic friction µk is defined for the case in which one surface
is sliding across another at constant speed. It is
kinetic friction force
coef f icient of kinetic f riction =
normal force

Ff
µk = (12)
FN

kinetic friction force = coefficient of kinetic friction × normal force

Ff = µk · FN (13)

7 Calculation of the resultant force for simple


physical systems
7.1 Motion of a box on a flat horizontal surface
The acceleration a of an object
P with mass m is dependent on the presence or
absence of a resultant force F in a specific direction of interest. For example,
if we consider the motion of the box shown in Figure 1, we are particularly
interested in the motion of the box parallel to and perpendicular to the flat
horizontal surface. Before we proceed any further, we chose to designate the
direction parallel to the flat surface as our x-direction and that which is per-
pendicular to the flat surface as our y-direction. Furthermore, the y-direction
above the flat surface is designated as the positive y-direction while that below
the surface is designated as the negative y-direction. Along the x-direction, we
assume that the box will tend to move to the right side hence this direction
becomes our positive x-direction and the direction opposite the motion of the
box i.e. the left side will become our negative x-direction.
With the issue of direction resolved, we can now shift our attention to
whether the box is motionless or moving in a particular direction. Naturally, a
box like the one shown in Figure 1 shows no tendency to change its velocity or
position along the y-direction when its lying on a flat surface, therefore, has no
acceleration along the y-direction i.e. ay = 0 m/s2 . This shows that the two
forces acting on the box in the y-direction i.e. FN and FW do not produce a
resultant force. Since FN points in a direction above the horizontal surface, we
take it to be a positive force. Similarly, the weight of the box FW which always
points downwards is taken to be a negative force. Therefore, we can get the
resultant of these two forces in the y-direction as
X
Fy = +FN + (−FW )

6
X
Fy = FN − FW (14)
With this expression of resultant force along the y-direction, we can now
write Newton’s Second Law of Motion as follows:
X
Fy = may
where
P
Fy is the resultant force along the y-direction

m is the mass of the box


ay is the acceleration of the box along the y-direction.
P
Substituting for the expression of Fy , we get

FN − FW = may (15)
Since the box shows no tendency to move along the y-direction, the accel-
eration is zero along this direction i.e. ay = 0 m/s2 . Substituting this value of
acceleration we get

FN − FW = m(0 m/s2 )

FN − FW = 0

FN = FW (16)
We therefore conclude that the normal force FN exerted by the flat surface on
the box is always equal to the weight of the box FW but acts in the opposite
direction. The flat horizontal surface whenever it “feels” the weight of the box
tends to push up against the surface of the box it is in contact with. This is a
consequence of Newton’s third law of motion.
Next, we shift our attention to the motion of the box along the x-direction.
A close inspection of Figure 1 shows that there are two forces acting on it along
the x-direction. The first force is the force F which could easily be a tensile
force if a rope is attached to the box. The second force is the friction force Ff
which always tends to oppose the motion of the box and thus always points in a
direction opposite the motion. With these two forces, we can get an expresion
for resultant force along the x-direction as follows:
X
Fx = +F + (−Ff )

X
Fx = F − Ff (17)

7
Along the x-direction, the box can behave in a number of ways. The box can
chose to remain motionless, move at constant velocity, move faster or slow down
depending on the presence or absence of a resultant force along the x-direction.
If the box chooses to remain motionless, then its velocity or position will not
change. In this case, the acceleration of the box along the x-direction will be
zero i.e. ax = 0 m/s2 . We write Newton’s second law of motion for a motionless
box as:
X
Fx = max (18)
Subsituting for the expression of resultant force and acceleration along the
x-direction, we get

F − Ff = m(0 m/s2 )

F − Ff = 0

F = Ff (19)
where
F is the force pulling the box
Ff is the static friction force.

Recall that the static friction force is given by

Ff = µs FN (20)
and that the normal force FN is equal to

FN = FW = mg (21)
Therefore, we get

F = µs mg (22)
If the box chooses to move with constant velocity, its velocity will remain
unchanged but its position will change. Also, in this case, the acceleration of
the box along the x-direction will be zero i.e. ax = 0 m/s2 . We write Newton’s
second law of motion for a box moving with constant velocity as:
X
Fx = max (23)
The expression of resultant force remains unchanged and acceleration along
the x-direction is zero. Substituting, we get

F − Ff = m(0 m/s2 )

8
F − Ff = 0

F = Ff (24)
where

F is the force pulling the box


Ff is the kinetic friction force.
Recall that the kinetic friction force is given by

Ff = µk FN (25)
and that the normal force FN is equal to

FN = FW = mg (26)
Therefore, we get

F = µk mg (27)
If the box chooses to move faster or slower, undergoes an acceleration ax
which has to be worked out. The acceleration is important because it connects
our equations of motion with the resultant force. In this case, we write Newton’s
laws of motion as
X
Fx = max (28)
Since the resultant force remains unchanged and the acceleration is not equal
to zero, we get

F − Ff = max (29)
Dividing both sides by the mass of the box, we get the acceleration in the
x-direction as
F − Ff
ax = (30)
m
If the pulling force F is greater than the friction force Ff , then the box moves
faster and faster as the acceleration in the x-direction is positive. Similarly, if
the pulling force F is less than the friction force Ff , then the box is slow down
to a halt.
If the pulling force is absent i.e. F = 0 N, then the acceleration of the box
alonng the x-direction is negative and will cause the box to come to a stop. This
process is called “sliding” and the acceleration is given by
Ff
ax = − (31)
m

9
7.2 Motion of a box supported by a rope
Let us now turn our attention to the motion of a mass m with weight FW
supported by a rope through a tensile force FT as shown in Figure 2. The mass
can only move along one direction i.e. up and down. We chose to designate this
direction as the y-direction. Since the mass supported by the rope will naturally
tend to move downwards if the rope was not present or got cut, we choose the
downward direction as our positive y-direction and the upward direction as
the negative y-direction. A short inspection of Figure 2 shows that only two
forces are acting on the mass. Since the weight of the mass FW is pointing in
the positive y-direction we designate it as a positive force. The tensile force FT
which “lives” inside the rope will be a negative force. Thus, we can get resultant
force on the mass along the y-direction as follows:
X
Fy = +FW + (−FT )

X
Fy = FW − FT (32)
Next, we can write Newton’s second law of motion for the mass supported
by the rope as follows:
X
Fy = may (33)
where
P
Fy is the resultant force on along the y-direction

m is the mass supported by the rope


ay is the acceleration of the mass along the y-direction
The mass supported by the rope can behave in a variety of ways. The
mass can chose to remain motionless, move downwards at constant velocity or
accelerate towards downards.
If the mass chooses to remain motionless, then the velocity and position
of the mass will remain unchanged. This shows that the mass undergoes zero
acceleration i.e. ay = 0 m/s2 . Substituting the expression for resultant force
and the value of acceleration, we get

FW − FT = m(0 m/s2 )

FW − FT = 0

FT = FW (34)
We conclude that the tensile force FT generated by the rope is equal in
magnitude to the weight of the supported mass FW .

10
If the mass supported by the rope is lowered towards the ground at constant
velocity, the position of the mass will change but its velocity will remain un-
changed. The acceleration of the mass along the y-direction will still be equal
to zero i.e. ay = 0 m/s2 . We can write Newton’s second law of motion for this
scenario as follows
X
Fy = may
Substituting for the resultant force and the acceleration, we obtain

FW − FT = m(0 m/s2 )

FW − FT = 0

FT = FW (35)
Again, we conclude that the tensile force FT generated by the rope as the mass is
lowered towards the ground is equal in magnitude to the weight of the supported
mass FW .
If the mass supported by the rope accelerates downwards or upwards, then
we can write Newtons second law of motion as
X
Fy = may
Substituting for the resultant force and the acceleration, we obtain

FW − FT = may

Making the acceleration ay the subject of the formula, we get


FW − FT
ay = (36)
m

11
The University of Zambia
School of Natural Sciences
Department of Physics
PHY 1010
Lecture 4B
Motion of a box on an inclined plane
Mr. Gift L. Sichone
April 27, 2021

Direction on an inclined plane


An inclined plane is a titled surface that makes an oblique angle (i.e. acute
or obtuse angle) with the plane of the horizontal. Consider the incline plane
shown in Fig. 1 that makes an angle θ with the horizontal. In this case, the x
and y-directions of the inclined plane and those of the horizontal surface are not
the same. For an inclined plane, we choose the x-direction to be the direction
parallel to the surface of incline plane. On the other hand, the y-direction is
chosen to be perpendicular or at 90◦ to the surface of the inclined plane.

Figure 1: An inclinced plane that makes an angle θ with the horizontal surface.

1
Box on an inclined plane
A box of mass m on an inclined plane exerts a weight FW = mg that is directed
downwards as shown in Fig. 2. The direction of the weight of the box FW = mg
lies neither in the x nor y-direction of the surface of the incline. Thus to find the
x and y-components of the weight FW along the x and y-direction of the inclined
plane, we have to resolve the weight FW . Fig. 2 also shows the resolution of the
weight of the box FW = mg into a x-component parallel to the surface of the
incline plane and a y-component perpendicular to the surface of the incline. As
can be seen from Fig. 2, the x-component of the weight of the box is equal to
FW sin θ = mg sin θ while the y-component is equal to FW cos θ = mg cos θ.

Figure 2: Resolution of weight FW of a box on an inclinced plane

Thus, a box of weight FW = mg tends to exert a force of FW cos θ = mg cos θ


perpendicular to the surface of the incline. As shown in Fig. 3, the incline is
“not happy” about the box exerting a force of mg cos θ on its surface, thus
according to Newton’s Third Law of Motion, the inclined plane exerts an equal
but opposite directed force on the box. This reaction force exerted by the incline
on the box is perpendicular to the surface of the incline and is referred to as
the normal force denoted FN . Thus, by choosing the upward direction to be
positive and the downward
P direction to be negative, we can write the net force
in the y-direction, Fy , as
X
Fy = FN + (−mg cos θ) (1)

X
Fy = FN − mg cos θ (2)
The box on the incline does not accelerate in the y direction i.e. the box
does not fly off from the inclined plane. Therefore, the acceleration of the box
in the y-direction ay is zero. We can therefore write Newton’s Second Law of
Motion for the box in the y-direction as
X
Fy = may = 0 (3)

FN − mg cos θ = 0 (4)

2
Figure 3: Reaction force from inclined plane on box.

Therefore, we get

FN = mg cos θ (5)
This equation FN = mg cos θ shows that the inclined plane only generates a
normal force FN that is equal in magnitude to the y-component of the weight
of box i.e. mg cos θ.
In the x-direction, the x-compoment of the weight of the box mg sin θ always
tends to pull the box downwards along the incline. This force is only countered
by the friction force denoted Ff between the incline plane and the box. Since
the box will naturally tend to slid down, the friction force tends to oppose this
downward motion and tries to pull the box upwards. If we choose the downward
direction in which the box will naturally tend to slide as the positive direction
and the upward direction as the negative direction, we can get the net force on
the box in the x-direction as
X
Fx = mg sin θ + (−Ff ) (6)

X
Fx = mg sin θ − Ff (7)
We know that the friction force Ff can be expressed in terms of the coefficient
of friction µ and the normal force FN as

Ff = µFN (8)

Therefore, the net force in the x-direction can also be expressed as


X
Fx = mg sin θ − µFN (9)
If the box remains motionless and does not slid off the incline, then the
acceleration of the box along the x-direction ax is zero. Thus , we can write
Newton’s Second Law of Motion in the x-direction as
X
Fx = max (10)

3
X
Fx = m(0 m/s2 ) = 0 (11)

mg sin θ − µs FN = 0 (12)
where µs denotes the coefficient of static friction.

mg sin θ = µs FN (13)
Recall that FN = mg cos θ, thus substituting for FN , we get

mg sin θ = µs mg cos θ (14)


from this expession, we can obtain the µs as
mg sin θ sin θ
µs = =
mg cos θ cos θ

µs = tan θ (15)
This angle θ is the maximum angle at which the box on an inclinced plane
remains motionless without sliding. It is called the angle of friction. The
angle of friction is equal to the arctangent of the coefficient of static friction
between the surface of the inclined plane and the box.

θ = tan−1 (µs ) = arctan(µs ) (16)


Similarly, if the box begins to slide and accelerates down the incline, then
the acceleration of the box along the x-direction ax is nonzero. Thus , we can
write Newton’s Second Law of Motion in the x-direction as
X
Fx = max (17)

mg sin θ − µk FN = max (18)


where µk denotes the coefficient of kinetic friction.
Again, recall that FN = mg cos θ, thus substituting for FN , we get

mg sin θ − µk mg cos θ = max (19)


from this expession, we the acceleration ax as
mg sin θ − µk mg cos θ
ax = (20)
m
mg sin θ µk mg cos θ
ax = − (21)
m m

ax = g sin θ − µk g cos θ (22)

4
Factoring out the g, we get the acceleration the the x-direction as

ax = g(sin θ − µk cos θ) (23)


if the inclined surface is frictionless, then µk = 0, we get the acceleration
along the x-direction as

ax = g sin θ (24)

If the angle of the incline is set at θ = 0 , then we get the acceleration along
the x-direction as
ax = g(sin 0◦ − µk cos 0◦ ) (25)

ax = −µk g (26)

If the angle of the incline is set at θ = 90 , then we get the acceleration
along the x-direction as

ax = g(sin 90◦ − µk cos 90◦ ) (27)

ax = g (28)

5
The University of Zambia
School of Natural Sciences
PHY 1010
Lecture
Work, Energy and Power
Mr. Gift L. Sichone
Phone : +260764036560
Email : [email protected]
May 29, 2020

Introduction
This topic introduces the concept of work, energy, the work-energy theorem,
and the conservation of energy and power.

Learning Outcomes
The student should be able to:

1. understand the concept of work in terms of the product of a force and


displacement in the direction of the force;
2. calculate the work done in a number of situations;
1
3. recall and apply the formula for kinetic energy KE = mv 2
2
4. derive, from the defining equation W = F s, the formula P EG = mgh for
gravitational potential energy changes near the Earths surface
5. recall and use the formula P EG = mgh for gravitational potential energy
changes near the Earths surface;
6. define power as work done per unit time and derive power as the product
of force and velocity.
W
7. solve problems using the relationships P = and P = F v.
t

1
8. recall and apply the principle of conservation of energy to simple examples.

Work
The Work done by a force F is defined as the product of the force and the
displacement s in the direction of the force. In other word, work is the transfer
of energy from one body to another by way of the action of a force applied over
a distance. The object on which the force is applied must over if any work is to
be done.
Consider a situation where a force F acts on a body and the body undergoes
a displacement s. The component of the force F in the direction of displace-
ment s is given by F cos θ where θ is the angle between the Force F and the
displacement s. Therefore, the work done by the force is F cos θ multiplied by
the displacement s.

W = (F cos θ)(s) = F s cos θ


Work is a scalar quantity. It only has size and has no direction.
The SI unit for work is the newton-meter called the Joule. 1 Joule of work is
the work done by a force of 1 N when it displaces an object 1 m in the direction
of the force.
If F and s are in the same direction, the angle θ between F and s is θ = 0◦ ,
we get

W = F s cos θ = F s cos 0◦ = F s
If the Force F is in the opposite direction to the displacement s e.g frictional
force, then the angle θ = 180◦ , we get

W = F s cos θ = F s cos 180◦ = −F s


This is negative work.

Example 1
A force of 3.0 N acts through a distance of 12 m in the same
direction of the force. Find the work done.
The work done W is given by

W = F s cos θ

where F is applied force, s is the displacement and θ is the angle between


F and s.

W = (3.0 N)(12 m) cos 0 = 36 J

2
Example 2
An object is being pulled along the ground by a 75 N force
directed at 28◦ above the horizontal. How much work does the
force do in pulling the object 8.0 m.
The work done W is given by

W = F s cos θ

where F is applied force, s is the displacement and θ is the angle between


F and s.

W = (75 N)(8.0 m) cos 28◦ = 530 J

Example 3
A 0.3 kg object slides 0.80 m along the horizontal table. How
much work is done overcoming friction between the object and
the table if the coefficient of friction is 0.20.
The displacement s = 0.80 m, mass of object m = 0.3 kg, coefficient of
friction µ = 0.20
Friction force Ff is given by

Ff = µFN

The normal force FN is given by

FN = mg = (0.3 kg)(9.8 m s−2 ) = 2.94 N

Therefore, we get friction force as

Ff = µFN = (0.20)(2.94 N) = 0.588 N

The friction force always acts in a direction that is opposite the motion.
The angle between the friction force Ff and the displacement s is 180◦ .
The work done W is given by

W = Ff s cos θ

where Ff is friction force, s is the displacement and θ is the angle between


F and s.

W = (0.588 N)(0.8 m) cos 180◦ = −0.47 J

3
Example 4
How much work is done against gravity in lifting a 3.0 kg object
through a distance of 0.4 m
To lift an object, an external force has to be applied on the object being
lifted. If we assume that the object is lifted at a constant velocity i.e the
object is not accelerating, then the pulling force will be equal to the weight
of the object.

Fpull = mg = (3 kg)(9.8 m s−2 ) = 29.4 N

The pulling force Fpull and the displacement s = 0.4 m are in the same
direction. Therefore, θ = 0◦ .
The work done W is given by

W = Fpull s cos θ

where Fpull is pulling force, s is the displacement and θ is the angle between
F and s.

W = (29.4 N)(0.4 m) cos 0◦ = 11.8 J

Energy
Energy is a measure of the change imparted on a object. Energy is given to an
object when a force does work on an object. The amount of energy transferred
to the object is equal to the work done. When an object does work, it losses an
amount of energy equal to the work done. Energy like work is a scalar and is
measured in Joules. An object that is capable of doing work possesss energy.

Kinetic Energy
Kinetic Energy (K.E) is the energy possessed by an object because of its motion.
If an object of mass m is moving with a speed v, it has translational energy KE
given by
1
KE = mv 2
2
where m is in kg and v in m s−1 , the units of KE are Joules.
Since a body with KE has velocity v and mass m, then the body has linear
momentum p.
p = mv
Squaring both sides of we obtain

p2 = (mv)2 = m2 v 2

4
Dividing both sides by m, we obtain
p2 m2 v 2
= = mv 2
m m
1
Multiplying both sides by , we obtain
2
1 p2 1
= mv 2
2m 2
Thus, we see a relationship between linear momentum p and kinetic energy KE
mv 2 p2
KE = =
2 2m

Gravitational Potential Energy (GPE)


GPE is the energy possessed by an object because of its position in a gravita-
tional field. A gravitational field is a region in which the force of attraction due
to gravity is felt. If an object with mass m falls through a vertical distance h,
it does an amount of work equal to mgh. The GPE is always given relative to
the Earth’s surface. If an object is at a height h above the surface of the Earth,
then its GPE is given by
P EG = mgh
where g is the accleration due to gravity, m is mass in kg and h is height in m.
Notice that mg is the weight of the object FW . The units of GPE are Joules.

Example 5
A 2.0 kg mass falls through a distance of 4.0 m. (a) How much
work is done on it by gravitational force (b) How much gravita-
tional potential energy did it lose?
The object has mass m = 2.0 kg, thus has a weight FW given by

FW = mg = (2.0 kg)(9.8 m s−2 ) = 19.6 N

The gravitation force is the weight of the object. The work done W is
given by
W = FW s cos θ
where FW is the gravitaional force, s is the height and θ is the angle
between F and s.

W = (19.6 N)(4 m) cos 0◦ = 78.4 J

The gravitational potential energy GPE lost is given by

GP E = mgh = (2.0 kg)(9.8 m s−2 )(4.0 m) = 78.4 J

5
The Work-Energy Theorem
The Work-Energy Theorem is also known as the Principle of Work and Kinetic
Energy. It states that when work is done on rigid body of mass m and there is
no change in GPE, the energy imparted on the rigid body can only appear as
KE.

Work done = change in KE

Example 6
A 0.5 kg block slides across a table with an initial velocity of
0.20 m s−1 and comes to rest at a distance of 0.70 m. Find the
average frictional force that retards its motion
Mass of block, m = 0.5 kg, displacement of block s = 0.70 m.
The initial velocity of block u = 0.20 m s−1 , therefore block had initial
Kinetic energy given by
1 1
KEi = mu2 = (0.5 kg)(0.20 m s−1 )2 = 0.01 J
2 2
When the block comes to rest, its final kinetic energy v = 0 m s−1 , there-
fore block has final kenetic energy given by
1 1
KEf = mv 2 = (0.5 kg)(0 m s−1 )2 = 0 J
2 2
Change in kinetic energy , ∆KE, is given by
∆KE = KEf − KEi = 0 J − 0.01 J = −0.01 J
The minus sign shows that the block lost kinetic energy.
We know use the principle of work and kinetic energy also known as the
work energy theorem, which states that the work done is equal to the
change in kinetic energy. Therefore, we obtain the work dones as
W = ∆KE = −0.01 J
But friction force Ff is always directed opposite the direction of motion.
Therefore, the angle between the friction force and the motion is θ = 180◦ .

W = Ff s cos θ = Ff (0.70 m) cos 180◦ = −0.01J

−Ff (0.70 m) = −0.01J


Dividing both side by -0.70 m, we obtain
−0.01 J
Ff = = 0.014 N
−0.7 m
The friction force Ff is found to be 0.014 N.

6
Example 7
The coefficient of friction between a 900 kg car and the pavement
is 0.80. If a car is moving at 25 m s−1 along the level pavement
when it begins to skid to stop, how far will it go before stopping?
The mass of the car m = 900 kg, the coefficient of friction µ = 0.80. The
initial velocity of the car is u = 25 m s−1 and since the car eventually
comes to a stop, its final velocity is v = 0 m s−1 . The change in kinetic
energy of the car is given by
1 1
∆KE = m(v 2 −u2 ) = (900 kg)((0 m s−1 )2 −(25 m s−1 )2 ) = −281 250 J
2 2
The friction force the car experiences as its skids is given by

Ff = µFN = µmg = 0.80(900 kg)(9.8 m s−2 ) = 7 056 N

The change in kinetic energy is equal to work done by friction force in


bring the car to a stop. The friction force is always directed opposite the
direction of motion. Thus by applying the work - energy theorem, we have

W = Ff s cos θ = (7 056 N)(s) cos 180◦ = −281 250 J

We obtain the displacement s as


−281 250 J
s= ≈ 40 m.
−7 056 N

The Principle of Conservation of Energy


The Principle of Conservation of Energy states that energy can neither be cre-
ated or destroyed, but is only transferred from one form to another. In other
words, the total energy of an object remains constant and is said to be conserved
over time. For an object with KE and GPE, the sum of its KE and GPE is
always constant.
K.E + GPE = constant

Example 8
A 1 kg projectile is shot upwards from earth with a speed of
20 m s−1 . How high is it when the speed is 8.0 m s−1 .
Mass of projectile, m = 1 kg. The initial velocity of projectile u =
20 m s−1 and the final velocity v = 8.0 m s−1 . The change in the ki-
netic energy of the projectile ∆KE is given by

1 1
∆KE = m(v 2 − u2 ) = (1 kg)((8.0 m s−1 )2 − (20 m s−1 )2 ) = −168 J
2 2

7
The negative sign in ∆KE shows that the projectile lost kinetic energy.
This lost kinetic energy according to the principle of conservation of energy
cannot be destroyed. Instead, it is only changed to a different from of
energy. In this case, the lost kinetic energy is changed to gravitation
potential energy. Thus, we have

∆GP E + ∆KE = 0

∆GP E + (−168 J) = 0

∆GP E − 168 J = 0

∆GP E = 168 J
The change in GPE is given by

∆GP E = mgh = (1 kg)(9.8 m s−2 )h = 168 J

we obtain the height h the projectile rises to as


168 J
h= = 17.1 m
(1 kg)(9.8 m s−2 )
The projectile rises to a height of 17.1 m.

Power
Power denoted P is defined as the rate of doing work. More, generally, Power
is how measure of how fast energy is transferred to an object.
Work done by force
Average Power =
time taken to do this work
W Fs s
P = = = F( ) = Fv
t t t
W
P = = Fv
t
The velocity v is in the same direction as the applied force F on the object.
The SI unit for Power is the Watt (W) and 1 W is equal to 1 J/s or 1 N m/s.
The other useful unit of Power is the kilowatt-hour(kW·h). If a force is
doing work at a rate of 1 kW ( i.e 1000 J/s), then in 1 hour, it will do 1 kW·h
of work. The kW·h is a measure of how much work a force does in 1 hour.

1 kW · h = 1000 J s−1 × 1 h = 1000 J s−1 × 3 600 s = 3 600000 J = 3.6 MJ

8
Example 9
How large a force is required to accelerate a 1 300 kg car from
rest to a speed of 20 m s−1 in distance of 80 m
The car has a mass m = 1300 kg and increases speed from rest i.e. u =
0 m s−1 to final velocity i.e. v = 20 m s−1 in a distance s = 80 m.
The change in kinetic energy ,∆KE, of the car is given by
1 1
∆KE = m(v 2 −u2 ) = (1300 kg)((20 m s−1 )2 −(0 m s−1 )2 ) = 260 000 J
2 2

Using the work-energy theorem, the work done by is equal to the change
in kinetic energy

W = F s cos θ = F (80 m) cos 0◦ = 260 000 J

The force provided by the car engine causing the acceleration is given by
260 000 J
F = = 3 250 N
80 m

Example 10
Calculate the average power required to raise a 150 kg drum to
a height of 20 m in a time of 60 s.
The weight of the drum, FW is given by

FW = mg = (150 kg)(9.8 9.8 m s−2 ) = 1 470 N

If we assume, that there is no acceleration when lifting the drum, i.e.


lifting force is equal to the weight of the drum, then our lifting force F
becomes
F = FW = 1 470 N
The work done in lifting the drum through a height of 20 m i.e. s = 20 m
is given by

W = F s cos θ = (1 470 N)(20 m) cos 0◦ = 29 400 J

The average power, P , is defined as the amount of work,W , divided by


the time to do the said work, t. We obtain the average power P as
W 29 400 J
P = = = 490 W
t 60 s

The average power required is 490 W.

9
The University of Zambia
School of Natural Sciences
Department of Physics
PHY 1010
Lecture 9
Density, Pressure and Buoyancy
Supplementary Study Materials
Mr. Gift L. Sichone
Phone : +260764036560
Email : [email protected]
August 19, 2020

Introduction
This lecture introduces the concepts of density and pressure. It also introduces
a special force called buoyant force that fluids always exert on objects. Arch-
inmedes’ principle of buoyancy is introduced and we show how it relates to the
buoyant force. We use Archimedes’s principle to explain why some objects sink
or float in a fluid.

Learning Outcomes
By the end of this lesson, the student should be able to:

1. define and use density;


2. define and use pressure;
3. derive the weight of a fluid;
4. derive, from the definitions of density and pressure, the equation P = ρgh;
5. use the equation P = ρgh;
6. understand the origin of the buoyant Force acting on a body in a fluid;

1
Table 1: The density of some substances at room temperature and pressure

substance ρ (kg m−3 )

water 1 000
seawater 1 030
oil 900
petrol 680
glycerine 1 260
ethanol 790
mercury 13 600
iron 7 900
gold 19 320
soldium chloride 2 160
aluminium 2 700
copper 8 920
magnesium 1 740
lead 11 340

7. state Archimede’s principle;

8. apply Archimede’s principle to solve simple problems involving sinking


and floating of objects;

Density
The density of a substance denoted ρ is defined as the mass of a substance in
a unit volume. A unit volume of any substance is a volume of cubic metre i.e.
1 m3 . In other words, density of a substance refers to the mass of substance
that can be contained in one cubic metre.
Mathematically, the density of a substance is defined as a ratio of mass of sub-
stance to the volume of substance and is given by
M
ρ=
V
where M is the mass of a substance in kg and V is the volume of a substance
in m3 . The SI unit for density is kg/m3 or kg m−3 .
Table 1 shows the densities of some common substances at room temperature
and pressure in kg m−3 . The density of a substance is an intrinsic property
i.e. does not depend on the size or shape of a substance. Properties such as
mass and volume that depend on size and shape are referred to as extrinsic
properties of substances.

2
Table 2: The specific gravity some substances

substance sp gr

seawater 1.03
oil 0.9
petrol 0.68
glycerine 1.26
ethanol 0.79
mercury 13.6
iron 7.9
gold 19.32
soldium chloride 2.16
aluminium 2.7
copper 8.92
magnesium 1.74
lead 11.34

Often, it is more convenient to express the volume of liquids in litres. The


litre is a metric unit of volume and is denoted as L or l. The cubic metre and
the litre are related as follows

1 m3 = 1 000 l

1
1l= m3 = 10−3 m3
1000
Sometimes, its is necessary to express the density of a substance in terms of a
dimensionless ratio called specific gravity. The specific gravity of substance
denoted sp gr is the ratio of the density of a substance to the ratio of the density
of some standard substance. This standard substance is usually water at 4◦ C
for liquids and solids, while for gases it is usually air.
ρ
sp gr =
ρstandard
Since sp gr is a dimensionless ratio has shown in Table 2, it has the same
value for all systems of units.

Density, Mass and Weight of Substances


Sometimes, the masses of substances such as liquids and gases can be expressed
more conveniently in terms of the density and volume of the substances under

3
investigation. Mathematically, the mass of a substance M can be given in terms
of the density ρ and volume V of the substance as

M = ρV

where ρ is the density of the substance and V is the volume of the substance.
Since the substance under investigation has a mass M and “feels” the grav-
itational pull of the Earth, we can obtain the weight of the substance FW as

FW = M g
where g = 9.8 m/s2 is the acceleration due to gravity on Earth.
Substituting for M = ρV , we get the weight of the substance FW as

FW = ρV g

where ρ is density of the substance, V is the volume of the substance and g is


the acceleration due to gravity.

Density and Pressure


Consider a beaker that is partially filled with a fluid to a height h. The beaker
has a base with a circular cross section area A. We can get the volume of the
fluid in the beaker V from the height of the fluid h and the cross section area
of the beaker A as
V = Ah

Figure 1: A fluid at rest in a partially filled beaker.

Next, we can get the mass of the fluid in the beaker M in terms of the
density of the fluid ρ and volume of the fluid V as

M = ρV
Substituting for V = Ah, we get the mass of the fluid in the beaker as

M = ρAh

4
Next, since the fluid in the beaker “feels” the pull of gravity due the Earth
i.e. fluids tends to pour downwards if the base of the beaker is opened, we can
get the weight of the fluid in the beaker FW as

FW = M g
Substituting for M = ρAh, we get the weight of the fluid in the beaker as

FW = ρAhg

Next, we seek to find the weight “felt” by a unit cross section area of the
base of the beaker. This ratio of the weight of the fluid to the cross section area
of the beaker or container is referred to as Pressure in Physics and is denoted
by P . We get the pressure P as
FW ρAhg
P = =
A A

P = ρgh
where ρ is the density of the fluid measured in kg/m3 , g = 9.8 m/s2 is the
acceleration due to gravity in and h is the height of the fluid column in metres.
Since the weight of the fluid FW is measured in Newtons (N) and the cross
sectional area of the beaker is measured in m2 , we get SI units of Pressure as
N/m2 . 1 N/m2 is also known as 1 Pascal (Pa).

1 Pa = 1 N m2
The idea of Pressure is not restricted to fluids only. Any substance that has
weight FW can exert Pressure on a surfaces underneath the substance.
Pressure is a vector. This is because to obtain Pressure, we have to divide
a force (i.e. a vector quantity) by the cross section area (i.e. scalar quantity).
Therefore, pressure has both magnitude and direction. The pressure and the
force point in the same direction. In the case where the force is the weight of
a substance, the pressure will be directed downwards towards the centre of the
Earth.
Therefore, we can more generally redefine pressure as ratio of the applied
Force F to the cross section area A. Mathematically, this idea of pressure can
be expressed as
F
P =
A
where F is the applied force in Newtons and A is the cross section area in
m2 .

5
Buoyant Force
Fill up an overflow can shown in Figure ?? with water. Next hang a spring
balance on a clamp as shown and attach a small solid metal bob to the end of
the spring balance. Take note of the weight of the metal bob from the reading
on the spring balance. Next, put the overflow can on the stand and slowly
submerge the bob into the water. You will notice that as you immerse the bob
in the overflow can, the level of water in the jug rises and even overflows into
the beaker where it is collected. The water which as been “evicted” from the
overflow can by the submerged bob is said to have been “displaced”.

Figure 2: A bob suspended in a fluid.

The volume of the displaced water collected in the beaker is related to the
volume of the solid bob submerged in the water. If the solid bob is completely
submerged, then the volume of the displaced water is equal to the volume of
the solid i.e.

Vw = Vbob
where Vw is the volume of the displaced water collected in the beaker and
Vbob is the volume of the solid bob submerged in the water.
We denote the density of water as ρw . Therefore, we get the mass of the
displaced water Mw as
Mw = ρw V w
Since the displaced water of mass Mw “feels” the pull of gravity due to the
Earth, we proceed to get the weight of the displaced water collected in the
beaker FW as
FW = M w g
Substituting for Mw , we get the weight FW as

6
FW = ρw gVw
In this case, where the solid bob is completely submerged, we also the weight
of the displaced water FW as

FW = ρw gVbob since Vw = Vbob


This weight of the displaced water FW = ρw gVw is known as the buoyant
force denoted FB . The buoyant force comes from the pressure exerted by
any fluid on solid objects that is either partially or completely immersed in the
fluid. Bacause the pressure increases with depth, the pressure at the bottom of
the object is always greater than at the top, hence the net force is an upward
force. Bouyant force is not always present in fluids. It is only present when an
object sinks or floats on a fluid.
To demostrate the existence of this buoyant force, take a small wooden block
and forcebly submerge it in jug filled with water.Immediately, after letting go
of the wooden block, the wooden block will be pushed out of the water and
accelerates towards the surface of the jug. The velocity of the wooden block
increases until it reaches the surface of the water. Since the wooden block
has mass and accelerates towards the surface of the jug, then there must be
force causing this acceleration accroding to Netwon’s second law of motion i.e.
FB = ma. This force is the buoyant force FB and it always tries to pushes
submerged objects out of any fluid. The greater the volume of the submerged
object, the more the bouyant force a fluid will generate to try to push out the
object from the fluid.

FB = ρ f g V s
where ρf is the density of the fluid, g is acceleration due to gravity and Vs
is the volume of the submerged solid.
The buoyant force FB is always an upward force and constantly counteracts
the weight of any object in a fluid. This is the reason why solid object submerged
in a fluid always “feels lighter” in a fluid. A liquid will only generate a buoyant
force FB once an object submerged in it. It is this buoyant force FB that enables
us to swin in water, that keeps wooden logs, boats and ships afloat on a river,
lake or at sea.

Archimedes’ Principle of Bouyancy


Archimedes’ Principle of Bouyancy states that the buoyant force FB that
is exerted on a body submerged in a fluid, whether fully or partially submerged,
is equal to the weight of the fluid displaced by that body.

FB = ρf Vf g
where the subscript f denotes the fluid.

7
If the weight of the fluid displaced by a body is less than the weight of the
object, the object will sink. Otherwise, the object will float, with the weight of
the displaced fluid being equal to the weight of the submerged object.

Figure 3: Why some submerged objects float, sink or remain motionless and
partially submerged.

Why Some Objects Float?


An object immersed in a fluid will float if the buoyant force FB is greater than
the weight of the object FW .
FB > FW
From Newton’s second law of motion, there is a net upward force that causes
the object to accelerate upwards. Taking the upward direction as positive and
the downward direction as negative, we get the net force Fnet as
X
Fnet = FB + (−FW ) = FB − FW > 0

where FB is the buoyant force and FW is the weight of the submerged solid.
The net force is positive and is directed upward. Therefore, the submerged solid
of mass ms accelerates upwards in the direction of the net force Fnet
X
Fnet = ms ay
where ms is the mass of the submerged object and ay is the acceleration of
the submerged object in the upward direction.

8
Why Some Objects Sink?
An object submerged in a fluid will sink if the buoyant force FB is less than the
weight of the submerged object FW .

FB < FW

From Newton’s second law of motion, there is a net downward force that
causes the object to accelerate downwards. Taking the upward direction as
positive and the downward direction as negative, we get the net force Fnet as
X
Fnet = FB + (−FW ) = FB − FW < 0
where FB is the buoyant force and FW is the weight of the submerged solid.
The net force is negative and is directed downwards.
Therefore, the submerged solid of mass ms accelerates downwards in the
direction of the net force.
X
Fnet = ms ay
where ms is the mass of the submerged object and ay is the acceleration of
the submerged object in the downward direction.

Why Some Objects Partially Sink and Remain Motionless?


An object will be partially immersed and motionless if the buoyant force FB is
equal to the weight of the submerged object FW .

FB = FW

Taking the upward direction as positive and the downward direction as neg-
ative, we get no net force Fnet .
X
Fnet = FB + (−FW ) = FB − FW = 0
where FB is the buoyant force and FW is the weight of the submerged solid.
The net force is negative and is directed downwards. From Newton’s second
law of motion, since the Fnet = 0, there submerged object does not experience
any acceleration i.e. ay = 0. It therefore, remains motionless and partially
submerged in the fluid. X
Fnet = ms ay = 0
where ms 6= 0 and ay = 0

9
The University of Zambia
School of Natural Sciences
Department of Physics
PHY 1010
Lecture 6
Angular Motion in a Plane
Mr. Gift L. Sichone
Phone : +260764036560
Email : [email protected]
June 4, 2020

Introduction to Angular or Circular Motion


Many useful everyday activities that make modern life meaningful require the
use and application of circular motion. Our cars once in a while have to pass
through a roundabout or bend so that we can change direction. Life on Earth
as we know it would not be possible, if Earth was not constantly orbiting the
Sun and spining on its own axis at the same time. The Moon orbits Earth once
every day. Our artifical satellites which bring television to our living rooms
are constantly orbiting the Earth in circular orbits. Our computers have hard
drives which spin on their own axis in order to read or store data. In hospitals,
circular motion is used in centrifuges and MRI scanning machines which are
used for diagnosing diseases. In our kitchens, we use mixers and blenders to
prepare food that keeps us health. Without mankind exploiting and harnessing
the powers of angular motion, life as we know it would not be possible.

Learning Outcomes
The student should be able to:
1. define and use angular displacement, angular velocity and angular
acceleration;
2. solve problems using equations of uniformly accelerated motion in a cir-
cular path;

1
3. describe qualitatively motion in a curved path due to a perpendicular
force, and understand the centripetal acceleration in the case of uniform
motion in a circle
v2
4. recall and use centripetal acceleration equations ac = and ac = rω 2
r
mv 2
5. recall and use centripetal force equations Fc = mrω 2 and Fc =
r

Description of Circular Motion


Angular Displacement
Consider a car going around a circular round about of length s and radius r.
The length of the round about s and the radius r are related as follow:

s = 2πr

where s is the linear displacement of the round about if it was stretched out
in a straight line and r is the radius of the round about. The SI units of both
s and r are meters. On the other hand, 2π is an angle measured in radians
and in circular motion, is referred to as angular displacement. The angular
displacement is denoted θ.
Thus, we can rewrite the above relation between s and r as

s = θr

Thus we get angular displacement θ in terms of linear displacement s and radius


of a circle r as
s
θ=
r
The SI units for θ are radians (rad in short).
The angular displacement θ can also be expressed in terms of degrees(◦ ) and
revolutions (rev in short).

1 rev = 360◦ = 2π rad.

Example 1
The bob of a pendulum 90 cm long swings through a 15 cm arc,
as shown in Fig. 1. Find the angle θ, in radians and in degrees,
through which it swings. We can get the angular displacement, θ, as
s 15 cm
θ= = = 0.17 rad
r 90 cm

2
Figure 1: A bob swinging from a simple pendulum

Angular velocity
The angular velocity denoted ω refers to how fast an object is moving round a
circular track like a roundabout or spining about its own axis. Angular velocity
ω is the rate of change angular displacement θ. Mathematically, the angular
velocity is given by
θ
ω=
t
where θ is angular displacement and t is the time. The SI units for angular
velocity are rad/s. If an object covers an agular displacement of θ = 2π in time
T , then T is referred to as the period. This measures how fast the object goes
round a circular path or spins. In Physics, this is reffered to as frequency and
is denoted f . The frequency f is given by
1
f=
T
and were T is the time required to complete one revolution. The period T has
SI units of seconds but frequency f is measured in Hertz (Hz). 1 Hz = 1 s−1 .
Therefore, if an objects covers an angular displacement of θ = 2π in time T ,
we can express our angular velocity ω as
θ 2π 1
ω= = = 2π
t T T

ω = 2πf
If θ of an object spining on its axis changes from initial angular displacement
θi to a final angular displacement θf in a time t, then the average angular velocity

3
ωav is given by
θf − θi
ωav =
t

Example 2
A fywheel turns at 480 revolutiopns per minute (rpm). Compute
the angular speed at any point on the wheel and the tangential
speed 30.0 cm from the center.
The angular speed ω in radians per second is
480 rev 480 rev × 2π rad/rev
ω = 480 rpm = = = 50.3 rad/s
1 min 60 s
The radius of the flywheel is r = 30.0 m = 0.30 m.
We can get the tangential velocity v as
v = rω = (0.30 m)50.3 rad/s = 15.09 m/s
v ≈ 15 m/s

Example 3
A fan turns at a rate of 900 rpm (i.e., rev/min).
(a) Find the angular speed of any point on one of the fan blades.
The angular speed or velocity ω in rad/s is
900 rev 900 rev × 2π rad/rev
ω = 900 rpm = = = 94.2 rad/s
1 min 60 s
(b) Find the tangential speed of the tip of a blade if the distance
from the center to the tip is 20.0 cm.
The tangential velocity or linear velocity v is related to ω by
v = rω
where r is distance from center of rotation in meters.
Therefore, we get v as
v = (0.20 m)(94.2 rad/s) = 18.8 m/s.

Angular Acceleration
The angular acceleration (α) of an object whose axis of rotation is fixed is the
rate at which its angular speed changes with time. If the angular speed changes
uniformly from ωi to ωf in a time t, then we get angular acceleration α as
ωf − ωi
α=
t
The SI units of α is rad/s2 .

4
Example 4
A wheel of 40 cm radius rotates on a stationary axle. It is
uniformly speeded up from rest to a speed of 900 rpm in a time
of 20 s. Find

(a) the constant angular acceleration α of the wheel


The radius of the wheel r = 40 cm = 0.40 m. The rest angular
velocity is ωi = 0 rad/s. The final angular velocity ωf in rad/s is

900 rev 900 rev × 2π rad/rev


ωf = 900 rpm = = = 94.2 rad/s
1 min 60 s
For a time t = 20 s, we get an angular acceleration of

ωf − ωi 94.2 rad/s − 0 rad/s


α= =
t 20 s
α = 4.71 rad/s2
(b) the tangential acceleration a of a point on its rim

a = rα = (0.40 m)(4.71 rad/s2 ) ≈ 1.9 m/s2


.

Deriving Equations for Uniformly Accelerated An-


gular Motion
There is a connection between linear physical quantities and angular counter-
parts. This connection extends even to the equation of motions (EOM) used to
describe circular motion. The EOMs for circular motion can easiy be derived
from the EOM for linear motion by using the following substitions

relation between s and θ


s = rθ

relationship between v and ω

v = rω

relationship between a and α

a = rα

5
Starting with the formula for linear average velocity vav given below
1
vav = (vf + vi )
2
Substituting for v = rω, we get
1
rωav = (rωf + rωi )
2
Cancelling out r on both side, we obtain
1
ωav = (ωi + ωf )
2
In linear motion, we known that the linear displacement s is related to the
linear average velocity vav and time t as follows

s = vav t
We can obtain an expression that shows how angular displacement θ relates to
average angular velocity ωav and time t, by substiting s = rθ and vav = rωav .
We obtain

rθ = rωav t
Cancelling out the r on both side, we get

θ = ωav t

Similarly, we can also obtain angular counterparts of the rest of the equations
of motion by making suitable substitutings.
For example, in the case of the following equation below, we substitute
v = rω and a = rα
vf = vi + at

rωf = rωi + rαt


Cancelling out the r on both side, we get

ωf = ωi + αt

Likewise, for the following linear equation below, we substitute s = rθ,


vi = rωi and a = rα. We get
1
s = vi t + at2
2

1
rθ = rωi t + rαt2
2

6
Table 1: Relations between Linear and Angular EOM

Linear Angular

1 1
vav = (vi + vf ) ωav = (ωi + ωf )
2 2

s = vav t θ = ωav t

vf = vi + at ωf = ωi + αt

1 1
s = vi t + at2 θ = ωi t + αt2
2 2

vf2 = vi2 + 2as ωf2 = ωi2 + 2αθ

Cancelling out the r on both side, we get


1
θ = ωi t + αt2
2
And finally, for the linear equation below, we substitute v = rω, a = rα and
s = rθ. We get

vf2 = vi2 + 2as

(rωf )2 = (rωi )2 + 2(rα)(rθ)


Expanding and simplifying, we get

r2 ωf2 = r2 ωi2 + 2r2 αθ


Cancelling out the r2 on both side, we get

ωf2 = ωi2 + 2αθ


We have seen that each equation of linear motion, has an structurally iden-
tical counterpart in circular motion as again show in Table

Example 5
A belt passes over a wheel of radius 25 cm, as shown in Fig. 2.
If a point on the belt has a speed of 5.0 m/s, how fast is the
wheel turning?
The angular velocity ω is a measure of how fast an object is turning. It is
related to the tangential velocity v and radius of the wheel r by
v = rω

7
Figure 2: A fly wheel turned by belt

Therefore, we can get ω in turns of v and r as


v
ω=
r

5.0 m/s
ω= = 20 rad/s
0.25 m

Example 6
A pulley of 5.0 cm radius, on a motor, is turning at 30 rev/s and
slows down uniformly to 20 rev/s in 2.0 s. Calculate
(a) the angular acceleration of the motor,
initial angular velocity ωi = 30 rev/s = 188.5 rad/s
final angular velocity ωf = 20 rev/s = 125.7 rad/s
time taken t = 2.0 s
therefore, we get the angular acceleration α as
ωf − ωi
α=
t

125.7 rad/s − 188.5 rad/s


α=
2.0 s

α = −3.14 rad/s2
(b) the number of revolutions it makes in this time, and
to get the number of revolutions, we first need to find the angular
displacement θ and then divide it by 2π rad since each revolution is
equal to an angular displacement of 2π rad. We get θ as
1
θ = ωi t + αt2
2

1
θ = (188.5 rad/s)(2.0 s) + (−3.14 rad/s2 )(2.0 s)2
2

8
θ = 370.7 rad

θ 370.7 rad
Number of revolutions = = = 59 rev
2π rad/rev 2π rad/rev

(c) the length of belt it winds in this time.


The linear displacement s is related to θ by

s = rθ

where r is the radius of the pulley in meters.

s = (0.05 m)(370.7 rad/s) ≈ 18.5 m

Example 7
A car has wheels of radius 30 cm. It starts from rest and accel-
erates uniformly to a speed of 15 m/s in a time of 8.0 s. Find the
angular acceleration of its wheels and the number of rotations
one wheel makes in this time.
Initial linear velocity u = 0 m/s
Final linear velocity v = 15 m/s
Time taken t = 8.0 s
We get the linear acceleration a as
v−u
a=
t

15 m/s − 0 m/s
a= = 1.9 m/s2
8.0 s
The tangential acceleration a is related to angular acceleration α and
radius of wheel r is given by
a = rα

We get the angular acceleration a as


a
α=
r
1.9 m/s2
α= ≈ 6.3 rad/s2
0.30 m
The linear distance the wheel travels in this time is given by
1
s= (u + v)t
2

9
Therefore, we get s as
1
s= (0 m/s + 15 m/s)(8.0 s) = 60 m
2

The angular displacement θ covered by a car wheel of radius r = 0.30 m


from s = 60 m is

s
θ=
r

60 m
θ= = 200 rad
0.30 m
The number of revolutions is given by
θ
Number of revolutions =
2π rad/rev

200 rad
Number of revolutions = = 31.8 rev
2π rad/rev

Centripetal Acceleration
A point mass m moving with constant speed v around a circle of radius r is
undergoing acceleration. Although the magnitude of its linear velocity is not
changing, the direction of the velocity is continually changing. This change in
velocity gives rise to an acceleration ac of the mass, directed toward the center
of the circle. We call this acceleration the centripetal acceleration and its
magnitude is given by

v2
ac =
r
where v is the speed of the mass around the perimeter of the circle.
Recall that v = rω, thus we can also rewrite the centripetal acceleration
as

v2 (rω)2 r2 ω2
ac = = = = rω 2
r r r

v2
ac = = rω 2
r

10
Centripetal Force
Objects do not like moving in a circular path. Given a choice, any object prefers
to move in a straight line. For an object of mass m to keep moving in a circular
path of radius r with velocity v, it must be forced. The force required to keep
an object in a circular path is referred to as centripetal force and is given by

v2
Fc = mac = m = mrω 2
r
The centripetal force has many “faces”. For a car going round a roundabout,
the friction force provides the centripetal force. In this case, the centripetal
force is the friction force.
Fc = Ff
In the instance of an mass attached to the end of a rope and being pulled round
in a circular path by a child, the centripetal force is provided by the tension
generated in the cord.
Fc = FT
For planets orbiting the Sun or satellites orbiting our planet Earth, the
centripetal force is provided by the pull of gravity.

Fc = FG

Example 8
A 200 g object is tied to the end of a cord and whirled in a
horizontal circle of radius 1.20 m at a constant 3.0 rev/s. Assume
that the cord is horizontal, i.e., that gravity can be neglected.
Determine
(a) the acceleration of the object
The mass of object m = 200.0 g = 0.2 m.
The radius of circle r = 1.20 m.
The angular velocity ω in rad/s is

3.0 rev × 2π rad/rev


ω = 3.0 rev/s = = 18.8 rad/s
1s
The acceleration ac is given by

ac = rω 2 = (1.20 m)(18.8 rad/s)2 ≈ 424 m/s2 .

(b) the tension in the cord.


The tension FT in the cord provides the centripetal force. Thus the
tension is equal to the centripetal force

FT = Fc = mac = (0.20 kg)(424 m/s2 ) ≈ 85 N

11
Example 9
What is the maximum speed at which a car can round a curve
of 25 m radius on a level road if the coefficient of static friction
between the tires and road is 0.80?
A car going round a curve on a level road is kept in its circular motion by
a centripetal force Fc . In this case, this centripetal force is provided by
friction force Ff between between the car tires and the road.
The centripetal force Fc is given by
v2
Fc = m
r
where m is the mass of the object undergoing circular motion, v is the
tangential velocity in m/s and r is the radius of the circular road in meters.
This centripetal force Fc is provided by friction force Fc which is given by

Ff = µs FN
where µ is the coefficient of static friction µs and FN is the normal force.
But FN is also equal to mg, thus we get
Ff = µs mg

Therefore, we get

Fc = Ff
v2
m = µs mg
r
Cancelling out the m on both side, we get

v2
= µs g
r
Multiplying both side by r, we get
v 2 = µs gr

Taking the square root on both side, we get



v = µs gr

p
v= (0.8)(9.8 m/s2 )(25 m)

v ≈ 14 m/s
The maximum speed is 14 m/s.

12
Example 10
A spaceship orbits the Moon at a height of 20 000 m. Assuming
it to be subject only to the gravitational pull of the Moon, find
its speed and the time it takes for one orbit. For the Moon,
mm = 7.34 × 1022 kg and r = 1.738 × 106 m.
A spacecraft orbit is circular and has a radius equal to the radius of the
moon plus the height of the spacecraft above the moon. Thus we get the
radius of the spacecraft orbit R as

R=r+h

where r is the radius of the moon and h is the height of the spacecraft
above the moon.

R = r + s = 1.738 × 106 m + 20 000 m

R = 1.758 × 106 m

The spacecraft is kept in its circular orbit by a centripetal force Fc that


keeps pulling the spacecraft towards the center of the orbit. This cen-
tripetal force Fc is

v2
Fc = ms
R
where ms is the mass of the spacecraft, v is the tangential velocity of the
spacecraft and R is radius of the spacecraft circular orbit.
The centripetal force Fc is provided by the gravitational pull of the moon
on the spacecraft FG . This FG is given by
ms mm
FG = G
R2

where G = 6.674 × 10−11 m3 · kg−1 · s−2 is the gravitational constant, ms


is the mass of the spacecraft, mm is the mass of the moon, and R is the
radius of the spacecraft circular orbit.
Thus we have
Fc = FG

v2 ms mm
ms =G
R R2
Cancelling out the ms and R on both side , we get

mm
v2 = G
R

13
Taking the square root on both side, we end up with
r
mm
v= G
R
s
(6.674 × 10−11 m3 · kg−1 · s−2 )(7.34 × 1022 kg)
v=
1.758 × 106 m

v = 1 669 m/s

The spacecraft has a tangential speed of 1 669 m/s.


The average angular velocity of the spacecraft ω is
v
ω=
R

we get ω as
1 669 m/s
ω= = 9.5 × 10−4 rad/s
1.758 × 106 m
To obtain the time to complete one revolution also known as period T ,
we divide the angular displacement for one revolution θ = 2π rad by the
angular velocity ω of the spacecraft. We get the period T as
2π rad
T =
ω

2π rad
T = = 6618 s
9.5 × 10−4 rad/s

The period T in minutes is given by


6618 s
T = ≈ 110.3 min
60 s/min

T ≈ 110 min 18 s

The time it takes the spacecraft to complete one orbit is 110 minutes 18
seconds.

14
Figure 3: A bob attached to a roof through a rope

Example 11
As shown in Fig. 3, a ball B is fastened to one end of a 24 cm
string, and the other end is held fixed at point Q. The ball whirls
in the horizontal circle shown. Find the speed of the ball in its
circular path if the string makes an angle of 30◦ to the vertical.
The ball is kept moving in a circle of radius r = (0.24 m) sin 30◦ = 0.12 m
by a centripetal force Fc provided by the x-component of the tension FT .
The centripetal force Fc is given by

v2
Fc = m
r
where m is the mass of the ball, v is the tangential velocity of the ball and
r is the radius of the ball’s circular path.
As earlier stated, the centripetal force Fc is provided by the x-componebt
of the tension FT thus, we get

Fc = FT sin 30◦

As the ball is whirled round in a circe, it does not go up or down. There-


fore, the ball experiences no acceleration in the y-direction. The ball has
weight FW = mg which should pull the ball down. However, the y- com-
ponent of the tension FT i.e. FT cos 30◦ pull the ball upward. Applying
the first conditin of equilibrium, we obtain
X
Fy = 0

15
Choosing the upward direction to be positive and the downward direction
to be negative, we obtain

FT cos 30◦ − FW = 0

FT cos 30◦ = FW = mg

Therefore, we obtain the tension FT as


mg
FT =
cos 30◦

Substituting this expression of FT into the centripetal force provided by


the x-component of the tension FT , we obtain
mg
Fc = FT sin 30◦ = ( )(sin 30◦ ) = mg tan 30◦
cos 30◦

Therefore, we now get

v2
mg tan 30◦ = m
r
Cancelling out m on both sides, we obtain

v2
g tan 30◦ =
r
Multiplying both sides by r, we get

v 2 = gr tan 30◦

Taking the square root on both sides, we obtain


p
v= gr tan 30◦

p
v= (9.8 m/s2 )(0.12 m)(tan 30◦ )

v = 0.82 m/s
The speed of the ball is 0.82 m/s.

16
Figure 4: A car going round on a road banking

Example 12
A curve of radius 30 m is to be banked so that a car may make
the turn at a speed of 13 m/s without depending on friction.
What must be the slope of the curve (the banking angle)?
For a car to move along a banked curve of radius r = 30 m, there must
be a centripetal force Fc to keep the car along its circular path. This
centripetal force Fc is given by

v2
Fc = m
r

The centripetal force is provided by FN sin θ. Thus we have

Fc = FN sin θ

However, a closer look at the provided figure shows that

FN cos θ = mg

Dividing the equation for Fc by FN cos θ, we obtain

FN sin θ Fc
=
FN cos θ mg

Fc
tan θ =
mg

mv 2 1
tan θ = ×
r mg

v2
tan θ =
gr

17
(13 m/s)2
tan θ =
(9.8 m/s2 )(30 m)

tan θ = 0.57

θ = arctan(0.57) ≈ 30◦

The banking angle is 30◦ .

Example 13
As shown in Fig. 5, a cylindrical shell of inner radius r rotates at
angular speed ω. A wooden block rests against the inner surface
and rotates with it. If the coefficient of static friction between
block and surface is µs , how fast must the shell be rotating if the
block is not to slip and fall? Assume r = 150 cm and µs = 0.30.

Figure 5: A bob stuck rotating ring

The wooden block remains at rest as long as the the ring is rotating at
a particular value of ω. The wooden block is being kept in position by
friction force Ff between the wooden block and the ring. If the ring stops
rotating, the friction force Ff disappears and the wooden block falls under
the weight of the wooden block.
The centripetal force Fc keeping the wooden block in position is given by
v2
Fc = m = mrω 2
r
The centripetal force Fc is provided by friction force

Fc = Ff = µs FN

18
But the the normal force FN = mg , thus we get

Fc = µs mg

Therefore, we get
µs mg = mrω 2

Cancelling out m on both sides, we obtain

µs g = rω 2

Dividing both sides by r, we get

µs g
ω2 =
r
Taking the square root on both sides, we get
r
µs g
ω=
r

We get r
(0.30)(9.8 m/s2 )
ω= = 1.4 rad/s
1.5 m
The ring should be rotating at an angular velocity of 1.4 rad/s.

19
The University of Zambia
School of Natural Sciences
Department of Physics
PHY 1010
Lecture 9B
Introduction to Mechanical Properties of Matter
Supplementary Study Materials
Mr. Gift L. Sichone
August 23, 2020

1 Introduction
This lesson will introduce idea of deformation of solids and the types of defor-
mations a solid can undergo under the action of an applied force. Under this
lesson, you will also be introduced to the concept of stress and strain together
with the types of stresses and strains a solid can experience.

2 Learning Outcomes
By the end of this lesson, the student should be able to:

1. Define deformation of solids;


2. Distinguish between elastic and plastic deformation of solids;
3. Define and state the four types of stresses and strains that exit in elastic
solids;

3 Definition of Deformation of Solids


Have you ever stretched a spring or bent a piece of wire? If you done these or
similar activities, then you definitely changed the original size of these solids.
For instance, if you stretched a spring without damaging it, you increased its

1
length during this time you applied the stretching force. Once you removed
the stretching force you had applied on the spring, it returned to its original
shape and size. On the other hand, if you stretched the spring with a large force
such that it got damaged, then upon removal of the stretching force the spring
did not return exactly to its original shape and size. For a piece of wire, your
experience of bending it was different. Once you had bent the wire and given it
some form, the wire permanently keep this new form or shape and did not try
to get back to its original form. These activities highlighted in this paragraph
are excellent examples of deformation of solids.
Deformation of a solid refers to change in length, volume or shape of that
solid as a result of an applied force. The force you apply on a solid causing
it to deform is referred to as the deforming force. The amount of deformation
you can cause on a solid depends on a number of factors. First and foremost,
the amount of deformation depends on the magnitude (or size) of the force you
are applying on the solid. If you apply a large stretching force on a spring, the
spring will undergo a large increase in length and vice versa. The amount of
deformation will also depends on how the force is being applied on the solid. A
stretching force applied on a spring will increase its length while compressing
force will decrease its length. Furthermore, imagine trying to stretch spring and
a piece of rubber. How much deformation do you think you will cause for the
same size of stretching force? Of course, your stretch will be more for the piece
of rubber compared to the spring. This shows that the amount of deformation
also hugely depends on the kind of substance the solid is made of e.g. rubber
or steel.
There are two main types of deformations in solids. As you are aware from
the preceding paragraphs, there is deformation where a solid is deformed and
once the deforming force is removed, the solid returns to its original shape or
size. This kind of deformation is referred to as elastic deformation. Elastic
deformation refers to temporary (non-permanent) deformation of a solid that is
recoverable (or self-reversing) once the deforming force is removed. An example
of elastic deformation is stretching a spring without damaging it such that the
spring returns to its original size and shape upon removal of the stretching force.
If you have a solid that completely recovers its original shape and size once you
remove the deforming force, then that solid is said to be perfectly elastic. No
real solids are perfectly elastic.
There is another kind of deformation called plastic deformation. This is
the kind of deformation that occurs when you bend a wire. As you are aware,
the wire you have bent will permanently keep its newly acquired form or shape.
Plastic deformation refers to a permanent deformation of a solid without fracture
under the action of a sustained deforming force. If the solid you have deformed
completely keeps its altered shape and size then that solid is said to be perfectly
plastic. You also need to be aware that no real solid is perfectly plastic. As
we delve deeper, you always need to keep in mind that perfect elastic or plastic
deformation does not occur in real solids. Real solids will always undergo a
mixture of both elastic and plastic deformation.
As you are now aware, for deformation of a solid to occur, a deforming force

2
has to be present and act on the solid in some way. Thereafter, depending on
the nature of the solid you are considering, the solid will experience a change
in length, volume, shape or all. For your deforming force to cause an increase
in length or volume of the solid, it has to increase the relative distance between
atoms of the affected solid without causing the solid to fracture. For a change
of shape to occur, the deforming force you have applied has to displace some
atoms more relative to others. You also need to be aware that, before you
applied a deforming force on the solid, the solid is at equilibrium and has a
specific distance between the atoms (inter-atomic distance). The deforming
force you apply on the solid, increases this inter-atomic distance yet your solid
always prefers to have the equilibrium interatomic distance. To achieve its
desire of having an equilibrium interatomic distance, your solid will setup a
force internally (inside of itself) that will pull back the atoms making up the
solid in a direction opposite to the direction of your deforming force. This force
which is created internally in the solid to oppose the deforming force is referred
to as the restoring (or recovery) force.
According to Newton’s third law of motion : For every acting force, there
is an equal but opposite reaction force, the restoring force (reaction force) is
equal in magnitude to the deforming force (acting force) but acts in direction
opposite. Due to this restoring force created internally in a solid as a result
of your applied deforming force, solids will tend to regain their original length,
volume or shape. This tendency of solids you have deformed to regain their
original length, volume or shape after the removal of the deforming force is
called elasticity.

4 Stress and Strain


4.1 Stress
You are now aware that a solid subjected to a deforming force experiences a
force within its interior structure that tends to balance the deforming force and
tries to restore the solid to its original condition. This restoring force created
inside the solid will be distributed across various interior surfaces of the solid.
Consider the two rods shown in Figure 1 below. These two rods have the
same original length lo = 1 m, are made of the same substance but have different
radii 5 cm and 10 cm respectively. From the formula of cross section area A given
by A = πr2 , where r is the radius of the rod, you can calculate the cross section
areas of the rods to be 0.0079 m2 and 0.0314 m2 respectively. If you subject
each rod to a deforming force of 450 N , an overall restoring force of 450 N is
setup internally from our use of Newtons third law of motion. This, however,
does not give you a clear picture of how the rods are “feeling” internally as a
result of this restoring force of 450N .
For you to have a clear picture of how these rods are “feeling” due to this
restoring force of 450 N , you need to calculate the restoring force per unit cross
section area of the rods . For the rod with radius of 5 cm, this restoring force

3
Figure 1: Schematic illustrations of two rods with different radii (a) 5 cm (b)
10 cm subjected to deforming force of 450 N.

per unit cross section area is 56 962 N/m2 (i.e. 450 N divided by 0.0079 m2 ).
For the rod with radius of 10 cm, the same restoring force per unit cross section
area is 1 433 N/m2 (i.e. 450 N divided by 0.0314 m2 ). You can clearly see that
the rod of radius 5 cm experiences or “feels” a much greater “restoring force
per unit cross section area” compared to the rod with 10 cm. This internal
“restoring force per unit cross section area” is termed as stress.

4.1.1 Definition of Stress


Stress refers to the force F per unit cross sectional area A of a solid due to
the presence of a deforming force. The symbol for stress is the Greek letter σ
(lower-case sigma).
F
σ= (1)
A
where F is the deforming force in Netwons (N) and A is the cross section
area in m2 . The SI units for stress are N/m2 or Pascals (Pa).

1 Pa = 1 N/m2

4
4.2 Strain
From our discussion of stress, you are aware that this deforming force of 450 N
will subject these two rods to different values of deforming stresses, 56 962 N/m2
for the rod with radius 5 cm and 1 433 N/m2 for the rod with radius 10 cm.
This causes the rods to undergo differing amounts of deformations with the rod
subjected to a greater deforming stress expected to undergo a larger increase in
actual length. If the rod with radius of 5 cm undergoes an increase in length
of let’s say 15 cm and the rod with radius of 10 cm undergoes an increase in
length of say 2 cm, it will quickly become difficult for you to determine which
rod has experienced the greater deformation.
You can see that the rod with the radius of 5 cm experiences a larger increase
in length because its subjected to a larger deforming stress. By the same token,
you can also see that the rod with radius of 10 cm undergoes a smaller increase
in length because its subjected to a lesser deforming stress. You can clearly see
that the actual increase in length or actual amount of deformation is not a very
suitable measure of deformation of solids. But if you compare the amount of
deformation (e.g. increase in length) to the original dimensions (e.g. original
length) of the solid, then you obtain a ratio that is independent of the shape or
form of the solid. This ratio of amount of deformation to the original dimensions
of the solid is referred to as strain.

4.2.1 Definition of Strain


Strain refers to the ratio of the amount of deformation (increase in length,
change in volume or shape) of a solid to its original dimensions (length, volume
or shape) of the solid. The symbol for strain is the Greek letter  (lower-case
epsilon). If you consider any of the rods in Figure 1 the amount of deformation
will be the increase in length (extension) while the original dimensions of the
rod will be the rods original length lo . Therefore, for a rod or a wire strain can
be defined as extension per unit length and is given by
extension
strain =
original length
We will denote the extension (i.e. increase or decrease in length) as ∆l where
∆ is the Greek capital letter Delta and l is the length. The final length of the
rod after deformation is denoted as l while its original length is denoted as lo .
Therefore, strain can be written more compactly as
∆l l − lo
= = (2)
lo lo
Where ∆l = l − lo is the extension in metres, l is the final length of the rod
after being deformed and lo is the original length of the rod in metres. From
Equation (2), you can clearly see that strain has no units.

5
5 Types of Stresses and Strains
Depending on the form of the solid in your possession (i.e. a rod or a cube) and
how you apply the deforming force on the solid, you will obtain different types
of stresses and strains.
Depending on the orientation of the deforming force, the deforming force
you apply on a solid can cause it to come under tension (i.e. be stretched).
This kind of deforming force is referred to as a tensile force and causes tensile
stress and tensile strain in the solid. If the deforming force causes the solid
to come under compression, then the deforming force is a compressive force.
There are three main types of stresses you will consider in this lesson as you
study the nature of deformation of solids. There is tensile stress, bulk stress
and shear stress. These stresses will correspondingly induce tensile strain,
bulk strain and shear strain the affected solids.

Figure 2: Schematic illustration of (a) a rod under the action of a tensile force
(b) a rod under the action of a compressive force

5.1 Tensile Stress and Compressive Stress


Tensile stress is the kind of stress setup in solids (i.e., rod or wire) under
tension while compressive stress builds up in solids under compression. Both
the tensile and compressive forces acting on the solid (i.e. the rod) will deform
the solid along the direction of the tensile or compressive force. The cross
section area of the rod being acted upon will be perpendicular ( ⊥ ) to the
direction in which you are applying the tensile or compressive force F⊥ causing
the deformation. Tensile stress,σtensile , or compressive stress, σcompressive is
equal to the magnitude of the tensile or compressive force, F⊥ , per unit cross
section area, A .
tensile force
tensile stress =
cross section area
compressive force
compressive stress =
cross section area
F⊥
σtensile = σcompressive = (3)
A

6
Where F⊥ is the tensile or compressive force and A is the cross section area
of the rod. The SI units of tensile or compressive stress are N/m2 or Pa.

Figure 3: Schematic illustration of a rod under the the action of (a) a tensile
stress (b) a compressive stress

5.2 Tensile Stress and Compressive Strain


Consider a rod fixed at one end and acted upon on the other end by a deforming
force F⊥ which brings about a change in its length as shown in the Figure 4.
If your deforming force is a tensile force , the rod undergoes an extension (an
increase in length), otherwise a compressive force will cause the rod to undergo
a contraction (a decrease in length).
If the initial length of the rod before you extended or compressed it is lo and
its final length is l after its been extended or compressed, then the deformation
is the extension, ∆l , of the rod is given by

∆l = l − lo (4)
If the rod is under tension (i.e. being acted upon by a tensile force) as shown
in Figure 4 (a) , then it will be stretched such that , l > lo , and the extension
∆l will be positive ( i.e. the rod is elongated). On the other hand, if the rod is
under compression (i.e. being acted upon by a compressive force), as shown in
Figure 4 (b) then is will be compressed (shortened) such that , l < lo . In this
case the extension will be negative valued.

7
Figure 4: Schematic illustration of a rod undergoing (a) a tensile strain (b) a
compressive strain

5.2.1 Definition of Tensile Strain and Compressive Strain


Tensile strain or compressive strain is the ratio of extension , ∆l , to the original
length lo .
elongation
tensile strain =
original length
elongation
compressive strain =
original length
∆l l − lo l lo l
tensile = compressive = = = − = −1 (5)
lo lo lo lo lo
l
The ratio λ = is called the stretch ratio. λis greater than one for tensile
lo
strain and less than one for compressive strain.
Equation (5) can be expressed in terms of the stretch ratio as

tensile = compressive = λ − 1 (6)

5.3 Bulk Stress.


Consider the cube shown in Figure 5, this cube is subjected to compressive forces
F⊥ of equal magnitude which act perpendicularly on all eight sides of the cube.
These compressive forces F⊥ as can be seen in Figure 5 act perpendicularly to
each side of the cube. The compressive force F⊥ per unit area A exerted on each

8
surface of the cube is known as pressure. As this pressure is being exerted on
the cube, the interatomic distance between atoms making up the cube reduces.
This results in an internal force per unit area called bulk stress or volume
stress being set up in cube to oppose the effect of applied external pressure.
According to Newton’s third law, this bulk stress is equal but opposite to the
direction of the applied pressure. The applied pressure is an external force
per unit area while the bulk stress is the internal force per unit area. A solid
subjected to these external pressures will therefore undergo decrease in volume
without a distortion in its shape. However, once the externally applied pressure
is removed, the bulk stress will attempt to restore the solid to its original volume.

Figure 5: Schematic illustrations of cube under the action of equal compressive


forces which are perpendicular to the cubes surfaces.

5.3.1 Definition of Bulk Stress


Bulk Stress is the kind of stress setup inside a solid when compressive forces
F⊥ are applied equally and perpendicularly to all sides (or surfaces) of the solid.
Bulk stress denoted , σbulk , is given by
F⊥
σbulk = = ∆P (7)
A
Where A is the area on which the perpendicular force F⊥ is being applied, and
∆P is the change in externally applied pressure. Bulk stress has units of N/m2
or Pa.

9
5.4 Bulk Strain
If the cube in Figure 5 had an initial volume of Vo , then after you apply these
compressive forces F⊥ equally on all faces of the cube, its volume will decrease
to V .The cube retains its shape but just undergoes a decrease in volume due
to the externally applied pressure which is also equal to the bulk stress. This
change in volume, ∆V , is given by

∆V = V − Vo (8)
If V < Vo then ∆V < 0 and the solid undergoes compression. On the other
hand, if V > Vo , then ∆V > 0 and the solid undergoes expansion.

5.4.1 Definition of Bulk Strain


Bulk strain is defined as the ratio of the change in volume of a solid, ∆V , to
the initial volume Vo of the solid.
∆V V − Vo V Vo V
bulk = = = − = −1 (9)
Vo V Vo Vo Vo
Where ∆V is the change in volume of the solid, V is final volume of the
solid and Vo is the initial volume of the solid. Bulk strain has no units.

5.5 Shear Stress


Consider the cube shown in Figure 6, the face of this cube on which the side
CD lies is acted upon by a deforming force F|| applied tangentially (or parallel)
to this face of the cube. Each face of the cube has area A . This tangential
force F|| applied in such a manner causes layers of atoms in the cube to be
relatively displaced with respect to each other but the cube as a whole remains
in the same position. This kind of deformation referred to as shearing. During
shearing, the layers of atoms of the cube closer to the face CD experience the
most displacement while those close to AB undergo no deformation. This results
in the cube undergoing only a change in shape but its volume remains the same
during the time the tangential force F|| is applied on the face CD of the cube.
As has been depicted in Figure 6, the point C will move to C 0 while D will move
to D60 under the action of F|| .
The tangential force F|| per unit area A applied on the face CD of the cube
causes the cube to shear resulting in a change in shape. Like before, through
the action of Newton’s third law of motion, this shearing will be opposed by
the cube. A shearing stress equal the tangential force F|| per unit area A in
magnitude will be setup inside the cube. This shearing stress will to oppose the
shearing effects of the externally applied tangential force F|| per unit area A on
the cube once the tangential force F|| has been removed so that the cube can
regain its original shape.

10
Figure 6: Schematic illustration of a cube being acted upon by a tangential
force F|| .

5.5.1 Definition of Shear Stress


Shear stress is defined as the tangential force F|| per unit area A . Shear stress
is not the same as pressure since the tangential force F|| is not perpendicular to
the area A on which it is acting. The shear stress denoted σshear is given by
tangential force
shear stress =
cross section area
F||
σshear = (10)
A
The SI units of shear stress are N/m2 or Pa.

5.6 Shear Strain


When the tangential force F|| is applied on the cube shown in Figure 6 the
maximum deformation occurs when the point C moves to C 0 and D moves D0
on the surface CD . This length of CC 0 and DD0 is the same and is in the
direction of the applied tangential force F|| . This magnitude of CC 0 and DD0
will be denoted by ∆x while the length CD and AC will be denoted by l .
The shear strain can thus be defined as the ratio of the maximum length of
deformation ∆x along the direction of the applied tangential force F|| to the
perpendicular length l in the plane of application of tangential force F|| .

11
5.6.1 Definition of Shear Strain
Shear strain is the ratio of the maximum length of deformation ∆x to the
perpendicular length l in the plane of action of the force.
∆x
shear = = tan θ (11)
l
In this definition of shear strain in term of tangent, the angle θ is in radians.
The conversion factor for angles in degrees to radians is 360◦ = 2π rad. Given
any angle of θ in degrees you can convert it to radians as follows:

θ(rad) = θ(deg) ×
360◦
π
For |θ << | , then tan θ ' θ, and Equation(11) becomes
2
shear = tan θ ' θ (12)

12
The University of Zambia
School of Natural Sciences
Department of Physics
PHY 1010
Lecture Notes
Temperature and Kinetic Theory of Gases
Mr. Gift L. Sichone
Phone : +260764036560
Email : [email protected]
September 27, 2020

1 Temperature and Thermometers


Temperature is one of the seven fundamental physical quantities in Physics. In
a nutshell, temperature can be defined as the degreee of “hotness” or “coldness”
of an object. Temperature affects our day to day lives in many ways. Normally,
we are interested in the temperature of some specific object. The finite portion
of matter whose temperature we want to measure is referred to as the system.
Everything outside of this system that has a bearing on the behaviour of the
system is referred to as the surrounding. Any physical system can be defined
in terms of physical quantities related to the behaviour of the system such as
the amount of substance, temperature, pressure and volume.
The temperature of any system can be measured using a device called a
thermometer. Thermometers exploit various physical properties that depend
on temperature such as thermal expansion of liquids, temperature dependent
voltages at the junction of two different metals and the temperature dependence
of electric resistance. The most common type of thermometer is the Liquid in
Glass thermometer which exploits the thermal expansion of liquids. A liquid
usually mercury or a coloured alcohol is sealed in a thin glass tube that has a
bulb at one end which serves as a reserve of liquid. Because the liquid expands
as the temperature rises, the liquid level in the thin capillary tube rises as the
temperature increases. The glass also expands but much less than the liquid.
Figure ?? shows a typical liquid in glass thermometer filled with a coloured
alcohol.

1
Figure 1: A liquid in glass thermometer filled with coloured alcohol

Gas thermometers are another commonly used type of thermometer in


the physical sciences. As shown in Figure ??, gas thermometers consist of a
gas filled bulb that measure the temperature of their surrounding by measuring
the pressure exerted by the dilute gas enclosed in a constant volume bulb and
connected with thin tube to a mercury manometer. In such gas thermometers,
the change in temperature of the the gas enclosed in the bottle is proportional
to the pressure exerted by the gas. Unlike, liquid in glass thermometers, gas
thermometers can accurately measure temperatures over very wide intervals.

Figure 2: A constant volume gas thermometer immersed in a liquid bath whose


temperature is to be measured

Due to the advert of the coronavirus disease also known as Covid-19, it


has become increasingly fashionable and acceptable to constantly measure the
temperature of individuals especially at points of entry in airports, hospitals,
malls and other public buildings without the need to make contact with the
person whose temperature is being measured. The fast spread of the Covid-19
across the globe has necessiated and pushed to the fore the use of contactless
thermocouple thermometers like the one shown in Figure ?? . Other types
of thermometers which are used in industries are ordinary thermocouples and

2
resistance thermometers.

Figure 3: A typical contactless thermocouple thermometer used to measure


the body temperature at points of entry in airports, hospitals, malls and other
public buildings

2 Thermometers and the Zeroth Law of Ther-


modynamics
Thermometers give quantitative and precise values of temperature. The num-
bers and units assigned to temperature are completely arbitrarily and need an
arbitrarily chosen standard system that is easily reproducible called a fixed
point. To precisely calibrate ( i.e. to mark graduations) a thermometer,
generally two fixed points that are globally easily reproducible are required de-
pending on the type of thermometer.
We use thermometers to first check if objects have the same temperature
and secondly to obtain accurate readings of the temperature of objects of
interest. Every thermometer uses a fundamental law of Physics called the Ze-
roth Law of Thermodynamics. When you place a thermometer in intimate
or close contact with an object whose temperature you want to measure, the
thermometer soon reaches a steady or constant reading called the tempera-
ture of the object, and we say that the object and the thermometer are in
thermal equilibrium with each other. If you now place this thermometer
in intimate contact with another object, the thermometer will similarly reach
a steady reading which will be the temperature of this object and will be in
thermal equilibrium with this object.
Now, suppose the thermometer reads the same temperature for the two
objects when they are in intimate contact with each other. The temperatures

3
of these two objects will not change but will remain constant. We say that
the two objects are in thermal equilibrium with each other or are at the
same temperature. Objects that have the same temperature are in thermal
equilibrium with each other. This in a nutshell is the essence of the Zeroth Law
of Thermodynamics.
Formally, the Zeroth Law of Thermodynamics can be stated as follows:
Two objects that are in thermal equilibrium with a third object are
in thermal equilibrium with each other. In other words, objects that are
in thermal equilibrium with each other have the same temperature.

3 Temperature Scales and Thermometers


3.1 Liquid in Glass Thermometers
A liquid in glass thermometer consists of a glass bulb containing a liquid (mer-
cury or a colored alcohol) attached to a fine glass tube with a very thin bore
evacuated of air and sealed. The glass tube has a numbered scale used to pre-
cisely track the rise and fall of the liquid column with temperature. If the liquid
in the bulb of a liquid in glass thermometer expands, say by 1%, this 1% extra
volume is forced to climb the evacuated narrow tube to quite a bit of height
thereby magnifying the expansion of the fluid. If the liquid column in a liquid
in glass thermometer expands to the same height when it is placed in intimate
contact with two different objects of systems separated by a special wall called
an adiabatic wall, then the two objects are at the same temperature using
the Zeroth law of thermodynamics. A good experimental approximation of an
adiabatic wall is thick wood, concrete, absestos or styrofoam. The other
kind of wall that you will encounter during temperature studies is a diathermic
wall. A good example of a diathermic wall is a thin sheet of metal.
To precisely calibrate a liquid in glass thermometer so as to create a tem-
perature scale, two fixed points that are global and easily reproducible have to
be found and marked on the thermometer. Next, the region between these two
fixed points is then divided into some number of equal steps. The first fixed
point called the freezing point of water is obtained or found by dipping the
thermometer in a bucket containing a mixture of ice and water. This point is
also called the melting point of ice. The liquid in bulb of the thermometer
will expand to a certain height and this height is marked and postulated to be
0◦ C . This temperature reading of 0◦ C is completely arbitrarily.
Next, we need to obtain the second fixed point. This is done by putting water
on a stove, heating it up under standard atmospheric conditions and when the
water begins to bubble, boil and evaporate, we dip the thermometer in the pot
of boiling water. The liquid in the bulb of the liquid in glass thermometer
will now be forced to climb up the bore of the glass tube to a much greater
height and will finally settle at a certain height. This height is marked and
is postulated to be 100◦ C . Once again, this temperature reading of 100◦ C is
completely arbitrarily.

4
Next the region on the thermometer capillary glass tube between the 0◦ C
mark and 100◦ C mark is divided into 100 equal parts. This is postulated to be
the temperature anywhere between 0◦ C and 100◦ C. Thus if the liquid in glass
thermometer is put in intimate contact with an object or a system and the liq-
uid column in the bore climbs 79% of the way to the top, then the temperature
is read as 79◦ C. This temperature scale is referred to as the centigrade scale
because there are 100 equal divisions between the two fixed points. The centi-
grade temperature scale is also known as Celsius scale in honor of Swedish
scientist Anders Celsius who first proposed its use in 1742.
The German physicist Gabriel Fahrenheit originally proposed a sort of
centigrade scale which came to be known as the Fahrenheit scale. He refer-
enced 0◦ F (degrees Fahrenheit) as the freezing point of a saline solution and
98.6◦ F as the temperature of the human body. On the Fahrenheit scale, the
freezing point and boiling point of pure water at standard atmospheric pressure
(1 atm) are 32◦ F and 212◦ F respectively.
The problem with liquid in glass thermometers lies in the fact that different
liquids do not expand at the same rate. Various liquid in glass thermometers
when being calibrated will agree at the fixed points because they have been
rigged to do so but will not necessarily agree in the region between the fixed
points.

Figure 4: The Celsius and Fahrenheit scales are commonly used in liquid in
glass thermometers

5
3.2 Constant Volume Gas Thermometers
The constant volume gas thermometer overcomes the limitation of the liquid in
glass thermometer especially when it comes to the rate of expansions. All gases
regardless of their nature expand at the same rate. To build a constant volume
gas thermometer, take a very dilute fixed mass of gas of constant composition
and fill in in bulk of constant volume. Connect the bulb using thin flexible tubes
as shown In Figure to mercury manometer . The diluted the gas, the better it
works in a constant volume gas thermometer.
Next, we submerge the bulb in a liquid bath sitted on a tripod stand under
a bunsen burner which as been switched off. Once the system (i.e. gas) has
settled down and there will be thermal equilibrium and the gas in the bulb will
occupy a certain volume V and a corresponding pressure from the height of the
liquid columns in the manometer liquid columns.

Figure 5: Schematic graphs of P V against T three gases used to calibrate a


constant volume gas thermometer on Celsius scale using the triple point of
pure water and the boiling point of pure water as fixed points

To calibrate any constant volume gas thermometer, we put the bath in which
the bulb is submerged on different surfaces like a hot plate or stove whose
temperature is known from a standard method like using the mercury liquid in
gas thermometer. On each surface, P and V for a sample of a gas used in a
constant volume gas thermometer is measured precisely at a known temperature
T and the product P V is taken and plotted on a graph of P V against T . In
an ice and water mixture, P V is obtained and the temperature T is postulated
to be 0◦ C . Similarly, on a surface of boiling water P V is again obtained and
the temperature T is postulated to be 100◦ C .
These two calibration points when connected by a straight line and divided

6
Figure 6: Schematic graphs of P V against T of three gases used to calibrate a
constant volume gas thermometer on Celsius scale extrapolated beyond 0◦ C

equally along the length of the straight line lead to an equal increase in the
product P V . The product P V for a gas is better than the volume of a liquid
e.g. mercury or alcohol because it does not depend on the nature of the gas. If
different samples of gas are used in the bulb and their P V against T graphs are
obtained, the graphs produced are straight lines but with different slopes. The
constant volume gas thermometer not only agrees between the fixed points but
also agrees all the way in between the fixed points. All gases have the property
that if they are calibrated between 0◦ C and 100◦ C , they agree in between.
If any constant volume gas thermometer is cooled below 0◦ C , the product
P V will eventually vanish i.e., become zero at a temperature of −273.15◦ C . All
other gases making a gas thermometer when cooled or when their P V against T
graphs are extrapolated backwards vanish at a temperature of −273.15◦ C . The
pressure due to a gas P vanishes at this temperature. This temperature is called
Absolute Zero Temperature because it is the lowest possible temperature
and no further cooling is possible beyond this temperature.
The 0◦ C is based on human obsession with water and is tied to life on planet
Earth, its not universal and natural. For any gas, the product P V vanishing
at −273.15◦ C is both natural and universal. Thus, on the Kelvin scale, 0K
temperature is set at −273.15◦ C as the first fixed point. The second fixed point
used on the Kelvin scale is the triple point of water. Water and ice can
co-exist at 0◦ C while water and steam can coexist at 100◦ C . By varying the
P , T and V in a unique way, a standard fixed point at which pure water,
ice and steam coexist simultaneously is found. This point is referred to as the
triple point of water and is assigned an arbitrarily value of 273.16 Kelvins,
abbreviated 273.15K . The region between the absolute zero temperature mark
0K and the triple point of water 273.15K is divided into 273.15 equal parts and
the magnitude of each division is defined as 1 Kelvin abbreviated 1K.
Since the triple point of water is a better standard fixed point than the

7
Figure 7: A phase diagram of pure water showing the position of the triple point
of pure water

Figure 8: Schematic graphs of P V against T of three gases used to calibrate a


constant volume gas thermometer on the Kelvin scale

8
Figure 9: A schematic graph of P V against T of a gas used to calibrate a
constant volume gas thermometer on the Kelvin scale

melting point of ice, the Celsius temperature scale has its zero point set to
Celsius temperature of the triple point of water at 0.01◦ C . The Kelvin scale is
like the Celsius scale except that the zero point has been shifted to 273.15◦ C.
The conversion between the Celsius temperatures Tc and Kelvin tempera-
tures T is

TK = T◦ C + 273.15◦ C (1)

T◦ C = TK − 273.15 K (2)

4 The Ideal Gas Law


The graph of P V against T on the Kelvin scale shows a linear relationship that
can be expressed as follows:

P V = cT (3)
where

P is the pressure of the gas used the constant volume gas thermomter
V is the volume of the gas of the constant volume gas thermometer
T is the absolute temperature of the gas in Kelvins
c is a constant of proportionality.

9
But for a gas that is sufficiently diluted and from experiments, it is found
that P V and T are precisely related as follows:

P V = N kB T (4)
where

N is is the number of atoms of a given kind


kB = 1.4 × 10−23 J K−1 is a constant called Boltzmann’s constant.

Notice that both N and kB are huge numbers. There is a natural number
called Avogadro’s number denoted NA which defines a mole of a substance
as 6.022×1023 atoms/mole. Thus the number of moles denoted n in N atoms
is given by
N
n= (5)
NA
Making N the subject of the formula, we get N in terms of n and NA as

N = nNA (6)

Substituting Equation (6) back into Equation (4), we get

P V = nNA kB T (7)
Next, we define a constant called the gas constant denoted R as follows

R = NA kB (8)

R = 6.022 × 1023 atoms/mol × 1.4 × 10−23 J K−1

R = 8.314 J/mol K
Substituting Equation (8) into Equation (7), we get

P V = nRT (9)

Equation (9) is known as the ideal gas law, and gases that obey it are
called ideal gases. An ideal gas is a very dilute gas such that the atoms or
molecules in the gas are so far apart hence they do not feel any forces between
each other unless they collide. All gases that are far removed from conditions
under which they condense show nearly ideal behaviour.

10
5 Kinetic Theory of Gases
We know that a gas is composed of atoms or molecules of a substance (or a
mixture of substances) which are free to fill any volume that contains them. We
are interested in studying an ideal gas for which the following conditions hold:
(i) The atoms or molecules composing the gas are very small and take up
negligible volume compared their container’s volume V .

(ii) The small gas molecules are quite far apart from each other and their is no
significant force acting between the atoms or molecules except when they
collide with each other and the boundaries of the container. There collis-
sion are all assummed to be perfectly elastic i.e. the linear momentum
and kinetic energy of the colliding atoms are both conserved.

(iii) The gas molecules are in constant radom motion.

Figure 10: A model of the motion of ideal gas atoms according to the kinetic
theory of gases

Consider a cube with sides L×L×L which has an ideal gas inside. The ideal
gas inside the cube has some pressure P . If the faces of the cube are not nailed
down to each other, then they will just fly off. We want to find the value of this
pressure and get reasons why there is pressure inside the cube. We also want to
get the connection between the temperature of the gas T and the mechanical
properties of the ideal gas atoms or molecules.
The ideal gas atoms or molecules are constantly bouncing off the wall and
every time each one of them bounces of the wall, its linear momentum changes
because the direction in which it moves changes. The wall changes the linear
momentum of the atoms or molecules colliding with it. The atoms or molecules
exert some force F on the wall and the wall pushes back with the same force F
according to Newton’s Third Law of Motion.
We are interested in finding the value of the force F exerted by these atoms
or molecules on the wall. To do this , we exploit the definition of force F as

11
the rate of change of linear momentum. Starting with Newton’s Second Law of
Motion , we get the resulatnt force F as

F = ma (10)
where
F is the resultant force
m is the mass of the ideal gas atom or molecule

a is the linear acceleration of the the atom as it bounces of the wall


We know from our previous studies of motion that linear acceleration a is
given by
v−u
a= (11)
t
Substituting Equation (11) into Equation (10), we get
 
v−u mv − mu pf − pi ∆p
F =m = = = (12)
t t t t
If we suppose that the ideal gas in the cube has N atoms or molecules
randomly moving in the cube, each with its own direction colliding against the
wall, the problem of calculating the force F becomes complicated. We simplify
the problem as follows:
(i) We approximate that one third of the N atoms are moving from left to
right, another one third of the N atoms are moving up or down and the
final one third of the N atoms are moving in and out of the board. Our
simplification assumes that atoms or molecules are moving only in these
3 primary directions. In reality atoms or molecules move in all directions
N
and even rotate. In our simplification, atoms are assumed to be going
3
back and forth between the two walls of the cube.
(ii) Next, we assume that all the atoms have the same speed v. This is a gross
and crude description of the problem.
Now we take one atom which hits the wall of a cube with initial linear mo-
mentum −mv and bounces back with final linear momentum mv. The negative
sign in the initial linear momentum is because after colliding with the wall, the
direction of motion of the atom will be opposite. However, the magnitude of
the velocity will still be the same as we have assummed that all the atoms move
with a constant velocity v. We get the change in linear momentum aka impulse
of the atom after colliding with the wall as

∆p = pf − pi = mv − (−mv) = 2mv (13)

12
The atom experiences this change its linear momentum after travelling a
distance of 2L at constant speed v . Therefore, we get atom’s the time of travel
∆t as
2L
∆t = (14)
v
Next, we can calculate the average force exerted by one atom on the wall
during a single collision Fone as

∆p 2mv 2mv 2 mv 2
Fone = = = = (15)
∆t 2L 2L L
v
But there is a large number of atoms colliding with the wall. Thus even if
each atom collides with a wall after travelling a distance of 2L , the force exerted
by the atoms on th ewall appears to be steady or constant. The average force
N
F̄ due to all atoms colliding with a wall of area L2 is given by
3
N N mv 2 N mv 2
F̄ = × Fone = × = (16)
3 3 L 3L
N
Only atoms apply this average force F̄ on the wall. The other two
3
directions are parallel to the wall, so the atoms do not apply any force on those
walls. The average pressure P̄ due to an average force of F̄ acting on a wall of
area A = L2 is given by

N mv 2
 

P̄ 3L N mv 2
P̄ = = = (17)
A (L2 ) 3L3
Now, setting V = L3 , we get the average pressure P̄ exerted on the wall as

N mv 2
P = (18)
3V
Multiplying on both sides of Equation (18) by V , we get

N mv 2
P̄ V = (19)
3
Now, recall from the ideal gas law, we have

P V = nRT (20)
These two equations (19) and (20) are equivalent, so we can write them as

N mv 2
PV = = nRT (21)
3
Therefore, we can then proceed to write

13
N mv 2
= nRT (22)
3
Multiplying on both sides of Equation (22) by 3, we get

N mv 3 = 3nRT (23)
Dividing both sides of Equation (23) by 2, we get

N mv 2 3nRT
= (24)
2 2
N
Next recall that n = and R = NA kB , thus substituting into Equa-
NA
tion (24), we have
 
N
3 (NA kB ) T
N mv 2 NA 3N kB T
= = (25)
2 2 2
Cancelling out the N on both sides, we have

mv 2 3kB T
= (26)
2 2
From our previous studies on work, energy and power, you should be able to
mv 2
notice and recall that the expression refers to the average kinetic energy
2
of the atoms. Therefore we can write
2
¯ = mv = 3 kB T
KE (27)
2 2
Equation (27) shows that the average kinetic energy of the atoms in the cube
is related to the absolute temperature as follows

¯ = 3 kB T
KE (28)
2
Equation (28) is one very important result of the kinetic theory of gases.
It gives the meaning of gas temperature as a measure of the average kinetic
energy of the gas atoms or molecules. So we can now define the absolute
temperature as a measure of of the average kinetic energy of the atoms or
molecules in an ideal gas. Thus, at absolute zero denoted 0K, the ideal gas
atoms or molecules cease to have any kinetic energy. They become motionless.
It also important to notice that we could Equation (28) to explain the mean-
ing of thermal equilibrium. Recall that substances in thermal equilbrium
have the same temperature. Therefore, if two ideal gases are in thermal equilib-
rium with each other, then their average translational kinetic energy per atom
or molecule is the same in both.

14
The expression for average kinetic energy of the ideal gas atoms or molecules
can modified and rewritten as

mv 2 3
= kB T (29)
2 2
m 2 3
(vx + vy2 + vz2 ) = kB T (30)
2 2
where vx , vy and vz are the components of the velocity of atoms in the 3 direc-
tions are atoms were allowed to move.
1 1 1 1 1 1
mvx2 + mvy2 + mvz2 = kB T + kB T + kB T (31)
2 2 2 2 2 2
Therefore, we get the average kinetic energy of the ideal gas atoms in each
direction as

1 1
mvx2 = kB T (32)
2 2
1 2 1
mvy = kB T (33)
2 2
1 2 1
mvz = kB T (34)
2 2
Last but not least, we can obtain from Equation (29) a kind of average
velocity of the atoms called the root mean square speed denoted vrms
Cancelling the 2 on both sides, we get

mv 2 = 3kB T (35)
Next, dividing by m on both sides, we get
3kB T
v2 = (36)
m
Finally, taking the square root on both sides, we get the root mean square
speed as


r
2
3kB T
v= v = (37)
m
r
3kB T
vrms = (38)
m
The rms speed not the same as the usual average speed or mean speed. Rather,
it is the speed that an atom or molecule with average kinetic energy possesses.

15
The University of Zambia
School of Natural Sciences
Department of Physics
PHY 1010
Lecture Notes
Description of Thermodynamic Systems
Mr. Gift L. Sichone
Phone : +260764036560
Email : [email protected]
October 1, 2020

1 The state of a system


Consider an ideal gas sitting in a container with a movable piston as shown in
Figure ??. An ideal gas is a very dilute gas such that the atoms or molecules
in the gas are so far apart hence they do not feel any forces between each other
unless they collide. At equilibrium, the ideal gas in a container with a movable
piston has a definite value of volume V , pressure P and absolute temperature
T . These three measurable physical quantities in the case of an ideal gas are
related via the ideal gas law as follows

P V = nRT (1)
The ideal gas in the container makes up the system and the state of this
system is a particular situation where the system has specified or definite values
of P , V and T . The measurable physical quantities used to describe the state
of a system are referred to as macroscopic coordinates. Since the pressure
P , volume V and absolute temperature T of an ideal gas in a container are
related by Equation (1), only two independent macroscopic coordinates P and
V are needed to describe the state of an ideal gas. The absolute temperature T
is not an independent macroscopic coordinate of a gas as it depends on P and
V via the ideal gas law.
If the macroscopic coordinates P , V and T which are used to describe a
system change in any way, then the system is said to undergo a change of
state. When a system is not influenced in any way by its surroundings it is

1
Figure 1: Schematic illustration of an ideal gas filled container with a movable
piston weighed down by lead pallets

said to be isolated. In practice, a system will be affected by its surroundings.


The surrounding may exert forces on the system or provide contact between the
system and a body that is at some definite temperature.

2 Types of equilibrium and equation of state


A system that has no unbalanced forces in its interior or between the system
and its surrounding is said to be in a state of mechanical equilibrium. If a
system is not in a state of mechanical equilibrium, it will undergo a change of
state which will only cease when mechanical equilibrium is restored. A system
in a state of mechanical equilibrium that does not tend to undergo a chemical
reaction or transfer of matter from one part of the system to another is said to
be in a state of chemical equilibrium.
A system in mechanical and chemical equilibrium which does not undergo
a spontaneous change in its macroscopic coordinates when it is separated from
its surroundings by a diathermic wall (e.g., a thin metal wall) is said to be in a
state of thermal equilibrium. In thermal equilibrium, all parts of the system
have the same absolute temperature T and this temperature is the same as that
of the surroundings. When these conditions are not reached, the system will
undergo a change of state until thermal equilibrium is reached.
When a system has conditions for mechanical, chemical and thermal equi-
libriums, it is said to be in a state of thermodynamic equilibrium. In
thermodynamic equilibrium, the system has no tendency whatsoever to un-
dergo a change of state. States of thermodynamic equilibriums are the only
ones described using macroscopic coordinates such as P , V and T . These
macroscopic coordinates used to describe states of thermodynamic equilibrium
are called thermodynamic coordinates. When any of three types of equilib-
riums which constitute thermodynamic equilibrium are not satisfied, the system
is said to be in a non equilibrium state and this state cannot be described

2
using thermodynamic coordinates because the macroscopic coordinates used to
describe the system are still changing and have not yet become steady or con-
stant.
For any system in a state of thermodynamic equilibrium, an equation called
an equation of state exists for that particular system that connects the inde-
pendent and dependent thermodynamic coordinates used to describe the system.
In the case of a container with a movable piston filled with an ideal gas in a
state of thermodynamic equilibrium, the independent thermodynamic coordi-
nates are P and V while the dependent thermodynamic coordinate is T . The
equation of state for an ideal gas connecting P , V and T is the ideal gas law
given by Equation (1). Since P and V can adequately describe any state of an
ideal gas, it is no longer necessary to use a P V against T diagram previously
used to construct Kelvin temperature scale for a gas thermometer. Instead P V
diagrams will be used going forward to describe any state of an ideal gas.

3 Intermediate states, state variables and inter-


nal energy of ideal gas
If three lead pallets are added on top of a movable piston as shown in Figure ??,
the piston will suddenly push down on the gas filled container due to the weight
of the lead pallets and will bounce back and forth a few times then with time
settle in a new location. After a while, the system will attain thermodynamic
equilibrium and the thermodynamic coordinates P and V of the system will
cease to change. This equilibrium state of the gas which we will refer to as
state 1 can be depicted with a dot on the P V diagram located at (P1 , V1 ) as
shown in Figure ??.
When one lead pallet is suddenly pulled out from the top of the piston leaving
behind two lead pallets, the gas inside the container suddenly expands and the
piston shots up and bounce back and forth. After a while, the piston settles
at another new location and the gas attains a new state of thermodynamic
equilibrium we will refer to as state 2 located at the dot on (P2 , V2 ) on the P V
diagram. By continuing to remove the reminder of the lead pallets in similar
fashion, two other states of thermodynamic equilibrium denoted on the P V
diagram as state 3 located at (P3 , V3 ) and state 4 located at (P4 , V4 ) are
obtained.
In between each state of thermodynamic equilibrium dshown on the P V
diagram shown in Figure ??, the system is not in a state of thermodynamic
equilibrium. There are no definite values of P and V that can be associated
with the ideal gas as it moves from one equilibrium state to the next one.
Different parts of the ideal gas will have different values of pressure P , thus
cannot be depicted as states on the P V diagram. Only when the system has
settled down and attained definite values of pressure P and volume V can that
particular state be represented on a P V diagram.
To always keep the system on the P V diagram at all times as it transitions

3
Figure 2: P V diagram depicting the different thermodynamic states an ideal gas
in a container with a movable piston passes through as lead pallets are suddenly
removed.

from one state to the next, the process of removing the lead pallets has to be
done quasistatically. A quasistatic process is one that proceeds sufficiently
slowly such that the system is always at equilibrium. In the case of the three
lead pallets used in Figure ??, they will have to be pulled very slowly from the
piston such that the system always remains in equilibrium. This approach is
impossible and not practical at all. A more practical approach is to replace the
3 lead pallets with grains of lead that produce the same amount of pressure as
shown in Figure ?? .

Figure 3: A schematic illustration of a ideal gas in container with movable piston


weighed down by grains of lead

Now, removing the grains of lead one at a time from the piston, the system
will quickly settle down to a definite value of P and V . This new state of the

4
system will be much closer to the initial state. However, during each settling
down process, the system has no definite value of P and V . The system
only has definite values of P and V once it has settled and attained a state
of thermodynamic equilibrium. Continuing to remove more grains of lead and
allowing the system to settle, several new states called intermediate states
which previously could not be obtained with the three lead pallets are obtained
as depicted in the P V diagram in Figure. The equation of state for ideal gases
can be applied to each intermediate state derived from this quasistatic process.

Figure 4: P V diagram depicting intermediate states found when the lead


pallets are replaced with grains of leads.

If the size of the grains of lead is reduced further, more intermediate states
are obtained between the already found intermediate states shown in the PV
diagram in Figure ?? . Using calculus, as the size of the grains is reduced
such that it approaches zero or vanishes, a continuous line is obtained showing
the path taken by the quasistatic process as the system moved from its initial
thermodynamic state to a final thermodynamic stat as shown in Fig. 4.5.
If all the grains of lead removed are returned on the piston one grain at a
time and the system is allowed to settle down each time a grain is returned, the
system ideally should return to its initial state following the same path. This
is referred to as a reversible process. Most often, processes are not reversible
when they are done quasistatically unless the system is completely frictionless.
When a system is returned to a previous state, those thermodynamic coordinates
that are always the same in that state are called state variables. Examples of
state variables for a system of an ideal gas in a container with a movable piston
are pressure P , volume V and absolute temperature T . State variables have the
same values regardless of how the system reaches a particular thermodynamic
state.

5
Figure 5: P V diagram showing a continuous path taken by the system as it
moved from its initial state to its final state..

The system of a gas in a container with a movable piston also has another
important state variable called internal energy of gas denoted U . In every
thermodynamic state , the system has a well-defined value of internal energy.
The internal energy U of a system is the total energy of every type possessed
by the system. For a system of ideal gas containing N atoms or molecules that
is allowed to move with f degrees of freedom , the Internal energy U of such as
an ideal gas is given by

N mv 2 f f f
U= = N kB T = nRT = P V (2)
2 2 2 2
Every state on the P V diagram has a specific value of internal energy U
associated with it and the value of U does not depend on how the state is
reached. The internal energy U is a state variable. For a system composed of a
monoatomic gas allowed to move in space only in 3 different directions (i.e. left
to right, up and down , in and out of the page) without rotating or vibrating
we have f = 3. In this case, the internal energy U of the ideal gas is given by
3 3 3
U= N kB T = nRT = P V (3)
2 2 2
For a system composed of a diatomic gas whose molecules in move in 3
directions in space and has rotational and vibrational motion, the degrees of
freedom increase to five i.e f = 5. In this case, the internal energy U of the
ideal gas is given by
5 5 5
U= N kB T = nRT = P V (4)
2 2 2

6
Chapter 12

12.0 Thermodynamics

12.1 Introduction
Thermodynamics is primarily associated with heat and temperature and their relation to energy
and work. It was developed in a quest to improve the efficiency of the early steam engines. In
general terms the basic concern is the transformation of thermal energy (heat) into mechanical
energy. It follows that a device or a system that converts heat into mechanical energy is called a
heat engine. Devices or systems may be simple or complex. The measurable quantities used to
describe a system are normally temperature, pressure and volume. In addition quantities such as
internal energy, heat, work and entropy are also used. For example, n moles of gas which is in
equilibrium in closed container will have a definite temperature, pressure and volume, such a
system will be said to be in a thermodynamic state. Thus the variables describing a given
thermodynamic state are called state variables. The properties of a given thermodynamic state will
remain the same as long as the state variables have the same values. An additional important
quantity that characterizes a system is internal energy (U). The internal energy constitutes of the
sum of all the kinetic and potential energies possessed by its atoms or molecules. A system’s
temperature is an indicator of its internal energy.

12.2 Zeroth Law of Thermodynamics


When two systems are each in thermal equilibrium with a third, then they are also in thermal
equilibrium with each other. The latter statement is called the zeroth law of thermodynamics. For
example if the temperature of an object measured with a thermometer is found to be the same when
a second object’s temperature is measured, we can conclude that the objects are in thermal
equilibrium with each other.

12.3 First Law of Thermodynamics


In discussing thermodynamics laws we consider heat as a form energy which it must conserved
like its mechanical equivalent. If a system is in a given thermodynamic state then it can be stated
that it has a definite amount of internal energy. When heat flows into a system it can do two things,
one it can increase the internal energy of the system and secondly it can provide energy to the
system allowing it to do work to the surroundings. This law is an extension of the law of
conservation of energy that includes heat energy and internal energy of the system under
consideration. The first law of thermodynamics says that the amount of energy Q going into a
system is equal to the sum of the increase in internal energy ∆U and the external work done W.

Q  U  W 12.1

It follows that a system in a given state has a definite amount of internal energy. When heat flows
into a system two are likely to happen:
i) The internal energy of the system increases or
ii) The heat may provide energy for the work to be done the surroundings.

For example, if we have a system comprising of a gas enclosed by a movable piston in cylindrical
vessel with a closed end see figure 12.1, and then heat is added to the system. The expanding gas
will move the piston and thus the gas does work on the piston. The piston moves against an external
force. Suppose heat was removed from the enclosed gas we expect the piston to move inwards.
The latter situation can be equated to the surroundings doing work on the system, in which case
the work will be negative.

Figure 12.1 Gas tight moveable piston in a cylinder.

In general many other systems give us similar results, we can conclude that the heat added to the
system should be equal to the increase in internal U plus the external work done by the system
W. It follows that the first law of thermodynamics is a law of conservation of energy which includes
thermal energy. Convention has it that when heat flows into the system the quantity of energy Q
is positive and when the heat flows out of the system then quantity of energy Q is negative. For
example in figure 12.1 if the piston is pushed downwards by an outside force F the work done is
negative. We can make an analogy with the definition of work in mechanics which states that:

Work=Force  Displacement  cos

Where the angle θ is the angle between the force vector and the displacement vector. Suppose that
the piston is pushed by a force F downwards and it moves a distance  y the work done W is:

W  F y cos  F y cos0o  F y 12.2

This result is because cos0o is equal to one as the force vector and displacement vector are in the
same direction. In figure 12.2 the piston is displaced by an amount  y and if the cross section area
of the cylinder is A then the volume increase is the product of this area and  y .

V  Ay 12.3
The pressure generated by the force F (arising from the weight of the piston) is P  F A . This
equation can be written in terms of the force as F  PA and substituting in equation 12.2 we get
that work by the expanding gas is:

W  F y  PAy  PV 12.4

Figure 12.2 Gas expansion in a cylinder

During expansion the volume occupied by the enclosed gas increases and the work done is
positive, whereas when the gas is compressed there is volume reduction i.e. V is negative, hence
the work done is negative. In the case of applying constant pressure on the gas, the work done is
simply calculated from the equation:

W  PV 12.5

12.4 Constant Pressure


In the case where expansion takes place under constant pressure, namely, isobaric process the work
done can be calculated using equation 12.5. This process can be plotted on graph of pressure versus
volume (P-V diagram) as in figure 12.3.

Figure 12.3 Pressure-volume plot at constant pressure.

The expansion takes place at a constant pressure P and the volume increases from VA to VB then
the work done in this case is:

W  PV  P(VB  VA ) 12.6

It should be noted that in equation 12.6 the term P(VB– VA) is the area under the process path. In
general the area under the process path is equal to the work done. In the case where the process is
reversed (compression) the area remains the same but the work done and the work is taken to be
negative.
12.5 Non-constant Pressure
The work done during a non-constant pressure process can also be worked out by finding the area
under the process path. However, it is tedious for such a process as we may have to use some form
of crude integration of the area on the P-V diagram as illustrated in figure 12.4.

Figure 12.4 Pressure-volume plot for changing pressure.

Considering the small shaded strip in figure 12.4, we can assume that the pressure over V is
constant and the work done is:

W  PV

The total work done will be equal to the sum of several such strips taken over the interval VA to
VB. Taking much smaller strips will yield a more accurate answer of the work done from A to B.

Example 12.1

A gas undergoes a process in which it is expanded from point A to B as shown in Figure 12.5.
Calculate the work done by the gas.

Figure 12.5 PV diagram

Solution

The work done by the gas is equal to the area under the process path AB and is equal to the sum of
the area of the triangle and the rectangle in figure 12.5.
Area of triangle = ½ base x height

= 1
2  600 10 6
 200 10 6  m 3  400  200 kPa

= 40 J

Area of rectangle =  600 106  200 106  m3   200  0  kPa

= 80 J

Total area under process path = 40 J +80 J = 120 J

The work done for this thermodynamic process is 120 J.

12.6 Internal Energy of an Ideal Gas


It was earlier on alluded to that the system’s temperature is an indicator of its internal energy. The
internal energy of a gas is mostly assigned to the kinetic energy possessed by the atoms/molecules
of a gas. Thus we can see the effect of temperature on the internal energy of a monatomic gas from
equation 12.7. This shows that the internal energy of an ideal gas depends only on its temperature.

KE  N ( KE )  3 2 nRT 12.7

Where N is the number of molecules in a gas, n is the number moles, and R the gas constant. In
the case of a monatomic gas the energy goes into mainly translational motion and its internal
energy is assigned to the kinetic energy as expressed in equation 12.8.

U  KEtrans.  3 2 nRT 12.8

It follows that the change in internal energy of a monatomic gas can be calculated from equation
12.9.

U  3
2 nRT 12.9

It should be noted that for molecular gases the energy flowing into them is partitioned into other
forms of motion such as rotation and vibration. These gases will have higher internal energy when
at the same temperature as a monatomic gas. Equation 12.9 can be re-written in a more general
form as:

U  K ( 1 2 nRT ) 12.10

Where K is an integer greater than or equal to 3. In the case of monatomic gases K is equal to 3.

12.7 Heat Transfer and Specific Heat Capacity of an Ideal Gas


The calculation of the amount of heat transferred into or out of an ideal gas depends on the process
used. Two common processes used are the constant-volume process and the constant pressure
process. These processes are discussed with reference to the first law of thermodynamics which is
about the conservation of energy.

12.8 Constant–Volume Process


Recall, the first law of thermodynamics states that energy added to system goes into increasing the
internal energy of the system plus work done, if any, by the system equation 12.1. In doing work
we expect a volume change in the system. Thus in a constant volume process where there is no
volume change  V  0  the work done is zero and all the energy goes into increasing the internal
of the system. The first law of thermodynamic reduces to:

Q  U (Constant-volume)

Referring to equation 12.9 for a monatomic gas we say that:

Q  U  3
2 nRT 12.11

We know that the only quantity that relates Q and T is the specific heat capacity of a substance
in equation 11.21.

Q
c
mT

We can now define a new quantity the molar specific heat capacity C by replacing mass m in the
last equation with the mole n.

Q
C 12.12
nT

To differentiate between the processes a subscript is attached to molar specific heat capacity term
and for constant-volume process Cv is used. Substituting for Q in equation 12.12 term from
equation 12.11 we get for a monatomic gas the following result:
3 nRT 3
Cv  2
 2R 12.13
nT

The above result can be generalized for molecular gases by appropriately changing the preceding
fraction.

12.9 Constant-Pressure Process


The work done in the case of a constant-pressure is simply the product of the pressure and the
change in volume ( W  PV ). It follows that the first law of thermodynamics can re-written as

Q  U  W  U  PV 12.4
At constant pressure the ideal-gas law equation 11.16 can be stated as

PV  nRT

The quantity of heat involved becomes

Q  U  nRT (Constant pressure)

In a similar manner we defined Cv, the molar specific heat at constant pressure, Cp is defined as

Q U  PV U nRT
Cp    
nT nT nT nT

At constant volume the ideal-gas law reduces to Q  U , hence if we substitute for U the
equation above becomes

C p  Cv  R 12.5

This result tell us that the molar specific heat capacity at constant-pressure is always larger that
the molar specific heat capacity at constant volume. This follows from the fact that in this process
some heat goes into work. Recall the molar specific heat capacity at constant volume for a
monatomic gas gave us the result that Cv  3 2 R . It follows that if we substitute this value in equation
12.5 we get

Cp  32 R  R  5 2 R

C p  5 2 R (monatomic gas) 12.6

The ratio of the two processes gives an approximately constant value  . The value of this constant
for monatomic gases is 1.67 and for diatomic gases 1.40. The values of  for a few selected gases
are given in table 12.1

  CP C 12.7
V

Table 12.1 Ratio of Cp to Cv of Some Gases

Gas 

Monatomic

Helium (He) 1.67


Argon (Ar) 1.67

Krypton (Kr) 1.69

Diatomic

Hydrogen (H2) 1.41

Oxygen (O2) 1.40

Carbon Monoxide (CO2) 1.40

Polyatomic

Steam (H2O) 1.30

Sulphur Dioxide (SO2) 1.29

Carbon Dioxide (CO2) 1.30

12.10 Zero-Heat Transfer Process


In the zero-heat transfer process transfer no energy is moved into or out of the system and it is
called an adiabatic process. For this to occur the process must be thermally isolated. In practice
this is very difficult to attain, however, if a process is very fast then it can be considered to be
adiabatic as there is very little time for the transfer of thermal energy? For this process Q = 0 and
the first law of thermodynamics becomes

0  U  W or U  W (adiabatic)

This implies that work is done at the expense of internal energy of the system or that the internal
energy of the system must decrease.

12.11 PV Diagram Representation of the Thermodynamic Processes


The first two processes, namely, the Constant-pressure and Constant-volume are simple as these
are represented by straight lines, for example, as in figure 12.3. The other two processes are
complex as for them there is an inverse relationship between the pressure and the volume arising
out of the ideal gas law. It states that if the temperature T is held constant then PV is also constant.
So that for the isothermal process the equation of the isotherm is

constant
PV  constant or P  12.8
V

A plot for an isothermal process involving compression at constant temperature is shown in figure
12.6.
Figure 12.6 PV diagram for the isothermal process.

Note that work is done during is the isothermal compression that is given by

V 
W  nRT ln  f  12.9
 Vi 

Where T, is the temperature at which compression takes place, Vf and Vi are the final and initial
volumes respectively. The final volume during this process is less than the initial volume and we
see that this is consistent with the fact that equation 12.9 will give us a negative number as the
natural log of a number less than one is negative.

In the case of the last process the adiabatic process where work is done at the expense of internal
energy the equation of the adiabatic is

constant
PV   constant or P  12.10
V

Where γ is as defined in equation 12.7. A plot of this process is in figure 12.7.

Figure 12.7 PV diagram for the adiabatic process

It should be noted that for adiabatic compression the pressure rises very rapidly when compared
to the isothermal process.

Example 12.2

A sample of air    1.40  is slowly compressed from a pressure of 1 atmosphere to 3 atmospheres.


The original volume v1 of air is 15 litres and the temperature is 300 K. The temperature is constant
during compression (isothermal process). Later on the air is suddenly (adiabatically) expanded
back to its original pressure of 1 atmosphere.
i) Sketch a PV-diagram of these processes,
ii) find the final volume and temperature,
iii) find the U , Q and W for each process, and
iv) number of moles of the air.

Solution

i)

ii) We can find the initial volume air by considering the isothermal part (AB) on the PV-
diagram using the ideal gas law.
PAVA  PBVB
P  1
VB  VA  A   15    5litres
 PB  3
The final volume VC at C has to be calculated using the adiabatic portion for which:
PCVC  PBVB

1 1
P   PB  
 3 1.4
VC  VC  B  or VC  VB    5   10.96litres
 PC   PC  1
To get the temperature at C we use the ideal gas law equation 11.14 from which we get
that
PBVB PCVC

TB TC
 P V 
 
TC  TB  C   C   300 1 10.96  219.1K
 PB   VB 
3 5 
iii) For the isothermal process there is no change in temperature  T  0  and therefore there
is no change in internal energy  U  0  .
The work done is given by equation 12.9.
WAB  nRT ln  B

V
V 
A
A A  V   1.0110 15 10  ln 515  1664.4J
  P V ln  VB
 A
 5 3
 
We can get QAB from the first law of thermodynamics.
QAB  U  W  0  (1664.4)  1664.4J
For the adiabatic process QBC = 0 and the result from the first law of thermodynamics is
that WBC  U BC .
iv) To find the number of moles of air the ideal gas law equation 11.16 re-written as,
PAVA 1.0110 15 10 
5 3

n   0.61mol
RTA 8.314  300 
12.12 Second Law of Thermodynamics
The second law of thermodynamics is concerned with the conversion of heat energy into usable
form i.e. into mechanical energy. Heat energy can be obtained from a variety of sources such as
wood, coal, oil etc. In our earlier discussion it was pointed out that heat is governed by the random
motion of atoms/molecules of a substance. To convert heat into a usable form we need to extract
energy from these random motions of atoms/molecules. This law limits the choice of fuels for
engines in general. The second law of thermodynamics is anchored on the flow of heat energy in
a given system. In everyday situations we know that ice left in the sun will melt and the temperature
of the ice can never increase when left in such an environment. In other words the heat flows into
the ice and a reverse flow is physically impossible.

12.13 Statement of the Second Law of Thermodynamics


It states that heat flows spontaneously from a substance at a higher temperature to a substance at a
lower temperature and cannot flow spontaneously in the opposite direction. This law is best
illustrated in figure 12.8.

Figure 12.8

12.14 Entropy (Re-statement of the second law of thermodynamics)


The concept of entropy states that if an isolated system made up of many parts is allowed to
undergo spontaneous change, it changes in such a way that the disorder increases or, at best, does
not decrease. Usually entropy is interpreted in terms of order and disorder of a system. Irreversible
processes lead to increase in the entropy of the universe, they lead to energy degradation meaning
that part of the energy becomes unavailable to do work. Entropy S ( ΔS denotes change in entropy)
is a state variable whose change is defined for a reversible process at a temperature T in which a
quantity of energy Q is the heat absorbed.
Q
S  12.11
T

Equation 12.11 is a measure of the amount of energy which is unavailable to do work or a measure
of the disorder of a system. The quantity Q is the amount of heat added in a reversible manner to
a system at temperature T. It should be noted that reversible processes do not change the total
entropy of the universe.

Example 12.3

A 10 g mass of helium gas is expanded by the isothermal process at a temperature of -85 0C to 4


times its original volume. What is the change in entropy of the helium gas?

Solution

The process is isothermal so ΔT = 0 and ΔU = 0, the from the first law of thermodynamics we get
that Q = W and work for this process is

 V f   0.01  4
W  nRT ln     8314 188  ln    5417J .
 Vi   4  1

Q 5417
S    28.8J/K
T 188

Example 12.4

Find the change in entropy that occurs when 50 g of ice melts slowly at 0 0C. Consider that heat
flows into the ice and regard it as an isolated system.

Solution

The ice is melting at a constant temperature hence the process is isothermal taking place at 273 K.
The heat needed for this process can be calculated using equation for the heat of fusion.

Q  mH f   50g  80cal/g  4.184J/cal   16736J

Entropy can be calculated from equation 12.11.

Q 16736
S    61.3J/K
T 273

12.15 Heat Engine


A heat engine is any device that converts heat energy into mechanical energy. We are familiar with
various types of engines powered by fuels such petrol, diesel, kerosene and rarely these days by
steam. All of the above have one thing in common in that they all transform heat energy into
mechanical energy by employing a repetitive cycle. They employ a working substance that is
returned to its initial state at the end of each cycle. Figure 12.9 is a general representation of a heat
engine that operates between a hot reservoir at temperature Th and a cold reservoir at temperature
Tc. In each operation cycle the engine absorbs an amount of heat Qh from the hot reservoir. The
absorbed heat is partly used to perform work, W, and the remainder, Qc, is transferred to the cold
reservoir.

Figure 12.9Schematic heat engine.

A heat engine obeys the law of conservation of energy. It follows that the heat energy from the hot
reservoir is portioned between doing work and being exhausted to the cold reservoir. In this case
the cold reservoir is the surrounding area. Since a heat engine obeys the law of conservation of
energy we can apply to it the first law of thermodynamics equation 12.5.

Q  U  W or that

Qnet  Qh  Qc  W  U 12.12

Where W is the work output per cycle and U is the change in the internal energy. A complete
cycle results in no net change in internal energy so that U is zero and equation 12.12 reduces to

W  Qh  Qc 12.13

To find the efficiency of a heat engine we use the definition of efficiency which states that the
efficiency of a “machine” is the ratio of its output to its input energy.

Output
Efficiency= 12.14
Input energy

Substituting for quantities in equation 12.4 and using the letter e for efficiency we get

W Qh  Qc Q
e   1 c 12.15
Qh Qh Qh

The heat energy Qc exhausted to the surroundings is the amount of energy that does not take part
in doing work and is responsible for the inefficiency of a heat engine. If there was no exhausted
energy (i.e. Qc= 0) to the surroundings then the efficiency of an engine will be hundred percent. It
is well known that this is a physical impossibility. We can conclude that there are definite
efficiency limits for heat engines.

Example 12.5

A certain engine absorbs 2000 J of heat energy from the hot reservoir and exhausts 800 J to the
surroundings during each cycle of operation.

i) What is the efficiency of this engine?


ii) How much work does it perform during each cycle?

Solution

i) The efficiency can be calculated using equation 12.15


Qc 800
e  1  1
Qh 2000
 1  0.4  0.6
i.e. the engine is 60% efficient.

ii) The work performed during each cycle is equal to the difference between the heat input
and the exhausted heat.

W  Qh  Qc

W = 2000 J – 800 J = 1200 J

12.16 Carnot Engine


A Carnot engine is an idealized engine which sets the maximum efficiency that can be obtained
by any engine operating at the same temperature extremes. No machine can be considered as ideal
as some heat energy is lost during operation due to friction. The friction losses can be accounted
for by the law of conservation of energy. A Carnot engine consists of a gas enclosed in a cylinder
with a movable piston and its operation cycle consists of two isothermal processes and two
adiabatic processes. In order that we approach Carnot efficiency, the processes involved must be
reversible and involve no change in entropy. Such conditions cannot be obtained in reality as real
engine processes are not reversible and physical processes lead to increase in entropy. The
efficiency of this engine can be calculated using equation 12.15.

12.17 Carnot Cycle


A Carnot cycle is usually represented on a PV diagram as, as shown in figure 12.10. Work is done
by the engine during the two expansions, and work is done on the engine during the two
compressions. The net work done per cycle of operation is the area enclosed by the four processes.
Figure 12.10 Carnot cycle.

The isothermal processes in the diagram are from A to B and from C to D. During isothermal
expansion A to B at temperature TA heat Qin is absorbed from the reservoir and work is done. In
the adiabatic expansion BC there is no further heat input but work is done at the expense internal
energy and the temperature falls to TC. During the isothermal compression CD the engine
surroundings are at temperature TC that is lower than TA. In this compression heat Qout is released
and work is done. To complete the cycle an adiabatic compression DA returns the engine to its
original state. The work done on the gas (compression) raises the temperature to TA. The total work
done during a cycle is equal to the area enclosed within the curves. The efficiency of a Carnot
engine can be calculated from equation 12.15. We know that the net work done in a cycle is equal
to the net heat transfer (Qin – Qout) as the change in internal energy is zero.

It is a fact that the amount of heat Q transferred from a Carnot engine is directly proportional to
the absolute temperature T.

Q∝ T

Q
  constant i.e.
T

Qc Tc
 12.16
Qt Th

It follows that the efficiency of a Carnot engine can be stated in terms of temperature by
substituting 12.16 into 12.15. The temperature must be in Kelvin for the latter equation to give
correct results.

Tc
eCarnot  1  12.17
Th

Example 12.6
A Carnot engine whose hot reservoir is at 550 0C takes 5000 J of heat per cycle and exhausts 2600
J of heat to the cold reservoir (surroundings).

i) Find the temperature of the cold reservoir, and


ii) Calculate the engine’s efficiency by using the temperatures and the heat quantities.

Solution

i) We are given the following: Th = 550 0C = 823 K, Qh = 5000 J and Qc= 2600 J. We can
use equation 12.16 to get the temperature of the cold reservoir.
Qc Tc
 i.e.
Qt Th
Qc 2600
Tc  .Th  .823  428K
Qh 5000
ii) Using equation 12.17

Tc
eCarnot  1 
Th
428
eCarnot  1   0.48
823
Using equation 12.15
Qc 2600
e  1  1  0.48
Qh 5000

Note that in both cases we get the same value of the efficiency, however this only holds for
Carnot engine. Any other engine operating at the same temperature extremes will result in a
lower value of the efficiency.

12.18 Refrigerators and Heat Pumps


A refrigerator operates in reverse fashion to a heat engine to extract heat from a low temperature
reservoir and transfer it to the high temperature reservoir. The natural tendency of heat is to flow
from a hot region to colder region. The operation of the refrigerator is contrary to what the second
law of thermodynamics states that heat cannot flow from a cold region to hot region on its own
accord. However if energy is supplied then heat can be forced to flow from a low temperature
region to a high temperature region. Refrigerators, air conditioners and heat pumps are devices
that do precisely that. Figure 12.11 is a schematic representation of a refrigerator.
Figure 12.11 Refrigeration

The process in which work is done on the system to lower the temperature of a substance is known
as a refrigeration cycle. In this case the energy flow as shown in figure 12.11 is the reverse of a
heat engine. A refrigerator operates at temperatures Tc and Th where the work input W allows heat
at a low temperature Qc to be moved to the high temperature region of the system. According to
the first law of thermodynamics that is for energy to be conserved, the energy into the system must
be equal to the energy out of the system.

Energy in = Energy out

Qc  W  Qh 12.18

In order for the refrigeration cycle to be very effective (efficient), it depends on removing large
amounts of heat for very little input work.

In refrigeration we specify the effectiveness of the system in terms of the coefficient of


performance, COP in place of the efficiency. This quantity is a ratio of the heat removed from the
cold reservoir Qc to the work input W.

Qc
COP= 12.19
W

If we re-arrange equation 12.18 so that W = Qh - Qc and then we substitute this in equation 12.19
we get:

Qc
COPrefrigerator  12.20
Qh  Qc
Referring to equation 12.16 where the quantity of heat energy involved is taken to be directly
proportional to the absolute temperature in Kelvin, then we can re-state equation 12.20 in terms of
absolute temperature.

Tc
COPMax  12.21
Th  Tc

The latter equation implies that when the difference between Tc and Th is small, less work is needed
to extract heat from the cold reservoir that is exhausted into the hot reservoir. This equation also
gives the maximum coefficient of performance.

12.19 Heat Pumps


Heat pumps such as air-conditioners operate in a similar manner to refrigerators. These can also
be used to heat up the interior of buildings during the cold season in moderate climates, in designs
that have a mechanism of reversing the flow of energy between the reservoirs. In this case the heat
pump heats up the cold interior as opposed to cooling it. It follows that the coefficient of
performance is the ratio of heat Qh delivered into the interior to the work W needed to deliver it.

Qh
Heat pump COP= 12.22
W

Since the work done is W = Qh – Qc we can substitute the term on the right side of this equation
into equation 12.22.

Qh
COPHeat pump  12.23
Qh  Qc

We can in a similar manner to a refrigerator define the maximum COP by substituting with
temperatures of the reservoirs in equation 12.23. However, this is a theoretical COP that can never
be achieved in reality.

Example 12.7

What mass of water at 00 C can a freezer with a coefficient of performance (COP) of 4.5 change
into ice-cubes at 00 C with a work input of 3.0 106 J. The latent heat of fusion, Hf, is 335 kJ/kg.

Solution

Qc
From equation 12.19 COP= , we need to calculate the energy used by the freezer using equation
W
11.23.

Q  mH f
mH f m  335 103
COP= 
W 3.0 106

4.5  3.0  106


m  40.3 kg
335  103
The University of Zambia
School of Natural Sciences
Department of Physics
PHY 1010
Lecture 12B
First Law of Thermodynamics
Mr. Gift L. Sichone
October 2, 2020

1 First Law of Thermodynamics


The first law of thermodynamics states that the change in internal energy
∆U = U2 − U1 of a system as it moves from one state to another on a P V
diagram is equal to the difference between the heat supplied to (or removed
from) the system ∆Q and the work done by (or done on) the system ∆W

∆U = ∆Q − ∆W (1)
where
∆U is the change in the internal energy of the system

∆Q is the heat supplied to (or removed from) the system. If the heat is
supplied to the system, ∆Q is taken to be positive. If the heat is removed
from the system then ∆Q is taken to be negative.
∆W is the work done by (or done on) the system. If the system expands
and does work on the surrounding, then ∆W is taken to be positive. If
the system contracts and the surrounding does work on the system, then
∆W is taken to be negative.
The change in internal energy ∆U depends only on the initial and final states
of the system. For a system of an ideal gas in a cylinder with a movable piston,
the internal energy of the system can be changed by

(a) putting the system on top of a hot reservoir (e.g. hot plate),

1
(b) allowing the system to do work on its surroundings or doing work on the
system.

The internal energy of an ideal gas depends only on absolute temperature T


.

Figure 1: A schematic illustration of a ideal gas in cylinder with movable piston


on a hot plate

2 First law of thermodynamics and work done


by an ideal gas
Consider a system of an ideal gas in a cylinder with a movable piston allowed
to do work on its surroundings as shown in Figure ??. The movable piston has
cross section area A and exerts pressure P on the gas. The force exerted by the
ideal gas on the piston is F = P A. The system is put on a hot plate at absolute
temperature T , causing thermal energy to be added to the system in the form
of heat. The ideal gas expands and pushes the piston upwards by a distance dy
. The work done by the expanding ideal gas on its surroundings is given by

∆W = F dy = P Ady = P ∆V (2)

where
P is the pressure exerted by the ideal gas on the movable piston
∆V is the change in volume of the cylinder.
Substituting Equation (2) into Equation (1), the first law of thermodynamics
becomes

∆U = ∆Q − P ∆V (3)
If the ideal gas in the cylinder is put on a hot plate, heat is added to the
system. If the ideal gas expands, pushing on the movable piston up, the ideal

2
gas does work on its surroundings and the system loses energy. If the movable
piston pushes down on the ideal gas, work is done of the gas by its surroundings
and the system gains energy. The work done is positive if the ideal gas expands
and negative if the ideal gas is compressed.
If the movable piston is nailed down so that it cannot move while the cylinder
with ideal gas has been put on top of a hot plate, the term P ∆V in Equation (3)
vanishes. In such as case, the first law of thermodynamics reduces to

∆U = ∆Q (4)
Equations (4) shows that the heat supplied by the hot plate into the system
goes into changing the internal energy of the system.
If the system is thermally isolated so that no heat can flow in or out of the
system from or to the surrounding, then the heat supplied or removed from the
system ∆Q in Equation (3) vanishes. If the piston is free to move, you can either
have a volume increase or a volume decrease. In this case, where the system
is thermally isolated, the first law of thermodynamics as given by Equation (3)
reduces to

∆U = −P ∆V (5)
If the ideal gas in the cylinder with a movable piston expands, the change
in volume ∆V is positive and the work done by the system on its surroundings
is positive. However, this positive work results in a negative change in internal
energy ∆U . The ideal gas pays for doing positive work against the movable
piston through a loss of internal energy U .
On the other hand, if the ideal gas is compressed by a movable piston, the
change in volume ∆V is negative and the work done by the gas is negative. The
surrounding does work on the system resulting in a positive change in internal
energy ∆U . The ideal gas molecules colliding with the movabel piston during
compression causes the ideal gas to gain kinetic energy.

3 Work done by an ideal gas under isothermal


process
If an ideal gas in a cylinder with a movable piston is put on top of a hot
plate at absolute temperature T . The absolute temperature of the system will
remains constant while ideal gas undergoes quasistatic expansion. This is an
isothermal process i.e., a constant temperature process. The P V diagram
shown in Figure depicts an isotherm (i.e. a constant temperature line on a PV
diagram) at absolute temperature T
The total work done by the ideal gas during the expansion is equal the total
area under the P V diagram. without doing a lot of calculus, we give the work
done ∆W by the ideal gas during an isothermal process as

3
Figure 2: P V diagram for an isothermal expansion of an ideal on an isotherm
at absolute temperature T .

 
V1
∆W = nRT ln (6)
V2
where
n is the number of moles of the ideal gas
R is the gas constant
T is the absolute temperature of the ideal gas
V1 is the initial volume of the ideal gas
V2 is the final volume of the ideal gas after expansion or compression.
For an isothermal process, the ideal gas does not undergo a change in ab-
solute temperature T . Therefore, the internal energy U of the initial state and
final state which lie on the same isotherm is the same, thus there is no change
in internal energy i.e. ∆U = 0 The first law of thermodynamics given by Equa-
tion (3), for an isothermal process becomes

0 = ∆Q − P ∆V (7)
Taking the P ∆V in Equation (7) to the left hand side, we get

∆Q = P ∆V (8)
where
∆Q is the heat supplied into the system
P ∆V is the work done by the system on its surroundings during this
isothermal process.

4
Since the work done by an ideal gas on its surroundings during an isothermal
process is given by Equation (6), we substitute into Equation (8) and get
 
V1
∆Q = nRT ln (9)
V2
Equation (9) shows that the heat supplied into the system from the hot plate
during isothermal expansion is used to do work against the surrounding without
changing the internal energy of the system.

4 Work done by an ideal gas under isochoric


process
Under an isochoric process (i.e. constant volume process), the volume of the
system is not allowed to change because the piston is nailed down. The system
does not do any work on its surroundings nor can the surrounding do work on
the system. If the movable piston is clamped and the cylinder put on a hot
plate, the P ∆V term in the first law of thermodynamics given by Equation (3)
vanishes since ∆V = 0. The work done by an ideal gas under any isochoric
process is always zero. The first law of thermodynamics given by Equation (3)
reduces to

∆U = ∆Q (10)
The heat energy ∆Q supplied from the hot plate into the system goes into
changing the internal energy ∆U of the system. The absolute temperature T of
the system is raised.

5 Work done by an ideal gas under isobaric pro-


cess
The total work done during an isobaric process (i.e. constant pressure process)
as the system moves from an initial state at absolute temperature T1 to a final
state at absolute temperature T2 is the total area under the P V diagram. Since
the pressure P is constant, we get the work done by the gas from only the
change in volume of the ideal gas ∆V . If the ideal gas changes volume from
initial volume V1 to final volume V2 , we get the work done by the ideal gas on
its surrounding as

∆W = P ∆V = P (V2 − V1 ) (11)
where
P is the pressure the ideal gas exerts on the movable piston
V1 is the initial volume of the ideal gas

5
Figure 3: P V diagram for various processes : an isochoric, isobaric, isother-
mal and adiabatic processes

V2 is the final volume of the ideal gas


The change in internal energy ∆U of the ideal gas exerting a pressure P on
the movable piston and undergoing a change in volume from V1 to V2 is given
by
f f
∆U = P ∆V = P (V2 − V1 ) (12)
2 2
Therefore, the first law of thermodynamics given by Equation (3) becomes
f
P ∆V = ∆Q − P ∆V (13)
2
Collecting the terms in Equation (13) with P ∆V , we get
f
∆Q = P ∆V + P ∆V (14)
2
From Equation (14), we see that the heat supplied ∆Q into the system from
the hot plate during an isobaric procoess is used to do work by the ideal gas
on its surroundings and to increase the internal energy of the system.
Factoring out the P ∆V from Equation (14), we get
 
f
∆Q = + 1 P ∆V (15)
2

6
Simplying the factor in the brackets, we get
 
f +2
∆Q = P ∆V (16)
2
For a monoatomic gas , the degrees of freedom is f = 3, therefore we get
 
3+2 5
∆Q = P ∆V = P ∆V (17)
2 2
For a diatomic gas , the degrees of freedom is f = 5, therefore we get
 
5+2 7
∆Q = P ∆V = P ∆V (18)
2 2

6 Work done by an ideal gas under a cyclic pro-


cess
A cyclic process is a process that starts and ends in the same state on a
P V diagram. Since the initial and final state in a cyclic process are the same,
the change in internal energy ∆U = 0. For a cyclic process, the first law of
thermodynamics given by Equation (3) reduces to

0 = ∆Q − P ∆V (19)
Simplifying Equation (19), we get

∆Q = P ∆V (20)
The heat supplied ∆Q to the system in a cyclic process goes entirely into doing
work on the surroundings. The work done is equal to the area enclosed by the
path on the P V diagram.
The following convention is adopted for work done and heat supplied under
cyclic process. For a clockwise path, ∆Q and ∆W are positive. The system
absorbs heat from the hot plate and does work on its surroundings. For a
counterclockwise path, ∆Q and ∆W are negative, work is done on the system
and heat is removed from the system.
For a cyclic process composed of isothermal expansion process A → B,
isobaric compressive process B → Cand isochoric process C → A, the area
enclosed by the path shown in Figure is the work done by the ideal gas during
this cyclic process.
The work done during isothermal expansion process A → B is positive
since VB > VA and is given by
 
VB
∆WA→B = nRT ln (21)
VA
The work done under isobaric compressive process B → C is negative since
VC < VB and is given by

7
∆WB→C = P (VC − VB ) (22)
The work done during the isochoric process C → A is zero since ∆V and
there is no area under the C → Aprocess. Thus we have

∆WC→A = 0 (23)

Figure 4: P V diagram showing a cyclic process composed of isothermal ex-


pansion process, isobaric compression process and an isochoric process

The net work done by the system during the entire cyclic process from
A → B → C → A is positive and is given by

∆W = ∆WA→B + ∆WB→C + ∆WC→A (24)


 
VB
∆W = nRT ln + P (VC − VB ) + 0 (25)
VA
 
VB
∆W = nRT ln + P (VC − VB ) (26)
VA
An alternate cyclic path going in the opposite direction composed of iso-
choric process A → C, isobaric expansion C → B and isothermal compres-
sion process B → A . The work done during the isochoric process A → C is
zero since ∆V = 0 and there is no area under the A → C process. Thus we
have

∆WA→C = 0 (27)

8
The work done under isobaric expansion process C → B is positive since
VB > VC and is given by

∆WC→B = P (VB − VC ) (28)

The work done during isothermal compression B → A is negative since


VA < VB and is given by
 
VA
∆WB→A = nRT ln (29)
VB
The net work done by the system during the entire cyclic process from
A → B → C → A is positive and is given by

∆W = ∆WA→C + ∆WC→B + ∆WB→A (30)


 
VA
∆W = 0 + P (VB − VC ) + nRT ln (31)
VB
 
VA
∆W = P (VB − VC ) + nRT ln (32)
VB

9
The University of Zambia
School of Natural Sciences
Department of Physics
PHY 1010
Lecture 13
Simple Harmonic Motion and Forced Vibrations
Mr. Gift L. Sichone
Phone : +260 764036560
Email : [email protected]
October 10, 2020

1 Introduction
This lesson introduces you to Simple Harmonic Motion (SHM) or Simple Har-
monic Oscillations (SHO). SHM is of great importance in physics because it is
exhibited by many physical systems. This lesson will show you what SHM is
and pinpoint the causes SHM in the absence of friction for a mass attached to
a spring. The lesson will also show you how to obtain the position, velocity
and accceleration of a spring-mass system undergoing SHM in the absence of
friction.

2 Learning Outcomes
By the end of this lesson, the student should be able to:

1. Define simple harmonic motion;

2. Calculate the periodic time, frequency, and natural frequency, maximum


velocity of a spring-mass harmonic oscillator
3. Calculate the position, velocity and accceleration of a spring-mass har-
monic oscillator;

1
3 Introduction to Simple Harmonic Motion
SHM in physics is very important because it is exhibited by many physical
systems. SHM involves repetitive movement of an object back and forth about a
position of equilibrium. During SHM, the displacement of the oscillating object
about the position of equilibrium on one side is equal to the displacement on
the other side.

Figure 1: Schematic illustration of a spring-mass system on a frictionless surface.

One of the most common systems that exhibits SHM is a spring-mass system
consisting of a mass hanging from a spring attached to a roof or a mass attached
to a spring on a frictionless surface. Figure 1 depicts a mass attached to the
spring that has come to rest after a long while moving back and forth on a
frictionless surface. The spring- mass system remains motionless (stationary) in
this position and is said to be in equilibrium because there is no net force causing
it to move. If the spring-mass system remains undisturbed, it will continue to
occupy this equilibrium position and remain stationary.
Let us consider a spring-mass system on in Figure 2 below in which a mass m
is attached to a spring with spring constant k on a frictionless surface. Assume
that the spring obeys Hooke’s law and let it have a relaxed length or extension
of x = 0 at stable equilibrium. If the mass is pulled by a deforming force
F , it stretches the spring over a distance x . If the mass is suddenly released
after being pulled, the mass attached to the spring will execute back and forth
motions about the equilibrium point x = 0 (i.e. the mass-spring system executes
oscillatory motions). The deforming force F = kx causes a restoring force
F = −kx to be set up in the spring, causing the system to oscillate about
x = 0.
From Newton’s second law of motion, the motion of the oscillating spring-
mass system is being caused by the restoring force F = −kx. Therefore, we
can write Newton’s Second Law of Motion in terms of the mass attached to the
spring m, the acceleration of the mass a and the restoring force F = −kx as

F = ma (1)

2
Figure 2: Schematic illustration of (a) a relaxed spring-mass system and (b) a
stretched spring-mass system on a frictionless surface

−kx = ma (2)
Making the acceleration a the subject of the formula, we get
k
a=− x (3)
m
So now we have the acceleration of the mass a in terms of the spring constant
k and the mass m.
k
Next, we proceed to set = ω 2 so that we have
m
r
k
ω= (4)
m
This ω is called the natural frequency of the spring mass oscillator.
The natural frequency is related to the frequency f of a simple harmonic
oscillator as follows:

ω = 2πf (5)
where f is the frequency measured in Hertz (Hz).
Making f the subject of the formula, we get
ω
f= (6)

The periodic time τ of a simple harmonic oscillator is related to the fre-
quency f as follows:
1
τ= (7)
f

3
Substituting for f , we get τ as

τ= (8)
ω
Substituting for ω, we get the periodic time τ as

τ=r
k
m
r
m
τ = 2π (9)
k
Substituting Equation (4) into Equation (3), we get the acceleration of the
spring mass oscillator as

a = −ω 2 x (10)
Now, it is possible to write the acceleration a in terms of the displacement
x as a second order derivative of time t
∆x
v= (11)
∆t

∆2 x
 
∆v ∆ ∆x
a= = = (12)
∆t ∆t ∆t (∆t)2
So very very small changes in x and t, the acceleration a becomes

d2 x
a= (13)
dt2
Substituting back into the equation for acceleration we get,

d2 x
= −ωx (14)
dt2

d2 x
+ ωx = 0 (15)
dt2
This Equation (15) is something called a differential equation in Advanced
Mathematics. This is the kind of equation that describes all forms of simple
harmonic equation.
The solution to Equation (15) is the function x = f (t) and is given by

x = xo cos(ωt + ) (16)
where
x is the position or displacement of the spring - mass oscillator at time t
from x = 0

4
xo is the maximum displacement of the spring mass oscillator from stable
equilibrium x = 0 called amplitude.
ω is the natural frequency of the spring mass oscillator
 is called the phase angle. The phase angle depends on the initial
conditions of the spring mass oscillator system. Often the phase angle is
zero i.e  = 0.

x = xo cos(ωt) (17)

Equation (16) shows that the displacement x of a simple harmonic oscillator


changes with time following the sine or cosine and is not constant. The sine
function motion of the simple harmonic oscillator is also known as sinusoidal
motion.
If the spring - mass system is oscillating in the y-direction, then its motion
is given by

y = yo cos(ωt) (18)
The velocity of a simple harmonic oscillator at time t is obtained by differ-
entiating the extension x with respect to time t and is given by
dx d
v= = (xo cos(ωt))
dt dt

v = −ωxo sin(ωt) (19)


The acceleration of the simple harmonic oscillator at time t is obtained by
differentiating the velocity v with respect to time t and is given by
dv d
a= = (−ωxo sin(ωt))
dt dt

a = −ω 2 xo cos(ωt) (20)
From these above equations, we are specifically intrested in the maximum
value of displacement, velocity and acceleration. These occur when the sin(ωt)
and cos(ωt) are equal to 1.
Therefore, we can get the maximum displacement in the x or y-directions as

xmax = xo , ymax = yo (21)


The maximum velocity of the simple harmonic oscillator is given by

vmax = −ωxo (22)


The maximum acceleration of the simple harmonic oscillator is given by

amax = −ω 2 xo (23)

5
Example 1
A particular spring stretches 0.2 m when a 0.5 kg mass is hung from it. Suppose
that this mass is replaced by a 2.0 kg mass and the system is then vibrated hor-
izontally as shown in Figure by displacing the mass 0.40 m from its equlibrium
position and releasing it. Find

(i) the spring constant k


The spring constant k is obtained from Hooke’s Law using the deforming
force F⊥ and the extension x. We get

F⊥
k=
x
where

F⊥ = (0.5 kg) × (9.8 m/s2 ) = 4.9 N


and extension x = 0.2 m.
Therefore, we get spring constant k as
F⊥ 4.9 N
k= =
x 0.2 m

k = 24.5 N/m

(ii) the natural frequency ω of the spring mass oscillator.


The simple harmonic oscillations of the spring mass oscillator are caused
when the 2.0 kg mass is displaced from its equilibrium position and re-
leased. The spring- mass oscillator begins to go back and forth about its
equilibrium position. The natural frequency ω of the oscillator is given
r
k
ω=
m

s s
24.5 N/m 24.5 kg m s−2 /m p
ω= = = 12.25 / s2
2.0 kg 2.0 kg

ω = 3.5 rad/s

(iii) the frequency of vibration of the system


The frequency of an oscillator f can be obtained from the natural fre-
quency ω as follows

ω = 2πf

6
Making f the subject of the formula, we get

ω
f=

We get the frequency f as

3.5 rad/s
f= = 0.557 /s
2π rad

f ' 0.56 Hz

(iv) the periodic time τ of the spring-mass oscillator


The periodic time τ is given by

1
τ=
f
Substituting, we get

1
τ= ' 1.79 s
0.56 Hz
(v) Plot the the position of the spring-mass oscillator over a time period of
5 seconds.

Figure 3: Position of spring mass oscillator over a time of 5 seconds

7
(vi) the maximum position of the mass,
We get the position of the spring mass oscillator from
x = xo cos(ωt)

To get the maximum position xmax , we set cos(ωt) = 1. Therefore, we get


xmax = xo = 0.4 m

This value of xmax is matches the amplitude of the plot of position.


(vii) Plot the the velocity of the spring mass oscillator over a time period of
5 seconds.

Figure 4: Velocity of spring-mass oscillator over a time of 5 seconds

(viii) The maximum velocity of the mass,


The velocity of the spring-mass oscillator is given by
v = −ωxo sin(ωt)

To get the maximum velocity vmax , we set sin(ωt) = 1

vmax = −ωxo

vmax = −(3.5 rad/s)(0.4 m)

vmax = −1.4 m/s

8
(ix) Plot the the acceleration of the spring-mass oscillator over a time period
of 5 seconds.

Figure 5: Acceleration of spring-mass oscillator over a time of 5 seconds

(x) its maximum acceleration,


We get the acceleration from the equation

a = −ω 2 xo cos(ωt)

To get the maximum acceleration, we set cos(ωt) = 1. Therefore, we get


acceleration as

amax = −ω 2 xo

amax = −(3.5 rad/s)2 (0.4 m)

amax = −4.9 m/s2

(d) the position, velocity and acceleration at of the mass at t = 1.0 s


The position of the spring-mass oscillator after t = 1 s is

x = xo cos(ωt)

9
Figure 6: Motion of spring-mass oscillator over a time of 5 seconds

x(t = 1 s) = (0.4 m) cos( (3.5 rad/s) · (1 s) )

x(t = 1 s) = −0.37 m

The velocity of the spring-mass oscillator after t = 1 s is

v = −ωxo sin(ωt)

v(t = 1 s) = −(3.5 rad/s)(0.4 m) sin( (3.5 rad/s) · (1 s) )

v(t = 1 s) = 0.49 m/s

The acceleration of the spring-mass oscillator after t = 1 s is

a = −ω 2 xo cos(ωt)

a(t = 1 s) = −(3.5 rad/s)2 (0.4 m) cos( (3.5 rad/s) · (1 s) )

a(t = 1 s) = 4.59 m/s2

(e) its velocity and acceleration when x = 0.10 m.

10
4 The Energy of a Simple Harmonic Oscillator
When the spring is stretched to maximum, work is done on the spring and it
undergoes a tensile strain. The work done on the spring is stored as elastic
potential energy. When the mass is released and its starts to move, the elastic
potential energy stored in the spring is converted into kinetic energy of the
moving mass. The total elastic potential energy UT is given by
1 1 1
UT = F xo = (kxo )xo = kx2o (24)
2 2 2
This total elastic potential energy is converted in kinetic energy and some
elastic potential energy at x.

UT = U (x) + KE (25)

1 2 1 1
kx = kx2 + mv 2 (26)
2 o 2 2
1
We proceed to cancel out the from both sides and get
2
kx2o = kx2 + mv 2
Collecting the terms with k on one side we get

mv 2 = kx2o − kx2
Factor out the k on the RHS, we get

mv 2 = k(x2o − x2 )
Divide on both sides by m, we get
k 2
v2 = (x − x2 )
m o
Taking the square root on both sidem we get
r
k 2
v= (x − x2 )
m o
r
k p 2
v= (xo − x2 )
m
p
v = ω (x2o − x2 ) (27)

p
v=±ω (x2o − x2 ) (28)

11
5 Damped and Forced Vibrations
In any vibrating system, there is always some loss of energy due to friction.
As a result, a mass attached to the end of a spring vibrates with constantly
decreasing amplitude as time goes by. Figure 7 shows the position of a spring-
mass oscillator that is vibrating in the ideal cases with no friction.

Figure 7: Position of spring mass oscillator over a time of 5 seconds

In a more realistic situation, the motion of our spring-mass oscillator is


influencced by friction. if the spring-mass oscillator is hung from the roof and
vibrates in air, then the friction is provided by the air. If the spring- mass
oscillator is vibrating on a very smooth surface, then friction comes from the
interaction between the horizontal surface and the mass. In either case, the
amplitude of the spring-mass oscillator will be seen to decrease constantly with
time until the oscillator becomes motionless. We say such as system is damped
and, that in this case, the amplitude of of the vibrations damp down fairly
quickly.
When the friction forces are very large, the spring-mass oscillator does not
vibrate at all; instead it simply returns slowly to its equilibrium position as
shown in Figure. Such a system is said to be overdamped. This situation
exists, for example, when the mass at the end of the spring is immersed in a
very viscous fluid. The mass does not move beyond the equilibrium position in
such as case. When the friction forces are just large enough that the system
returns to the equilibrium position without overshooting it, we say that the
system is critically damped.
If any system is to vibrate for an extended time, energy must be added
continuously to replace the energy lost doing work against friction forces. For

12
Figure 8: Damped oscillations or vibrations over a time in the presence of little
friction

Figure 9: Types of damped vibrations in the presence of varying amounts of


frictions

13
example, to keep a child swinging at constant amplitude on a swing, you must
push the swing from time to time to add energy to the system ( i.e. the child
and swing). In addition, everyone knows that there is a right and wrong way
to push a swing if it is to swing high. You must push with the motion of a the
swing and not against it. Only in this way can energy be added to the system
effectively. if you push against the motion, you can stop the vibration, since
the vibrating object must then do work on you, the pushing agent. Pushing a
swing is an example of a driven vibration or forced vibration.
In a driven system, the vibration is usually sustained by a repetitive driven
force acting on the system. This force has a frequency f , which may or may
not be the same as the natural frequency of the vibrating system fo . When the
f = fo , the drving agent is most effective at adding energy to the system. At all
other frequencies, the driving force is not quite in step with the motion of the
system, so the action of the force is less effective in adding energy. The drving
force is most effective when the its frequency f is equal to the natural frequency
of the of the system fo . In this case, we say that the force is in resonance with
the system. We call the frequency fo the resonance frequency of the system.

14
The University of Zambia
School of Natural Sciences
Department of Physics
PHY 1010
Lecture 14
Waves
Mr. Gift L. Sichone
Phone : +260 764036560
Email : [email protected]
October 10, 2020

1 The Origin of Waves and Wave Terminology


Many vibrating objects act as sources of waves. For example, sound waves can
originate from a vibrating tuning fork (see Figure 1) or from a vibrating guitar
string (see Figure 2).

Figure 1: Schematic illustration of a vibrating tuning fork producing sound


waves

In this lesson, we will first focus our attention on the study of waves on a
string and later extend our study to waves on a spring. We consider a taut (or

1
Figure 2: Schematic illustration of a vibrating guitar strings producing sound
waves

tight) string tied at one end and held by a hand on the free end as shown in
Figure 3. Thereafter, a pulse (or disturbance) is sent down the string by the
sudden up and down motion of the hand holding the string. The pulse travels
with speed v along the string and carries energy down the string. When the
pulse reaches a given point on the string, it causes that portion of the string
to momentarily acquire both kinetic and elastic potential energy. This energy
was given to the pulse by the source (i.e. up and down motion of hand) that
initiated the pulse. The energy moves with speed v down the string along with
the pulse. The pulse also serves as a record of what the source that initiated
it did. If the pulse is moving with velocity v along the string then after time
period t from the source, the distance travelled by the pulse will be given by

x = vt (1)
where

x is the displacement covered by the pulse in time t


v is the velocity with which the pulse moves along the string

When the source of the pulse vibrates with simple harmonic motion as shown
in Figure 4, the up and down motion of the string is transimitted down the string
with speed v, which is referred to as the wave speed. As a result, the string
has a sinusoidal shape at any given instant and this sinusoidal pattern travels
with speed v. As it moves the pattern carries energy down the string obtained
from the source of vibration.
The top of the wave marked A, C and E are called wave crests. The lowest
parts of the wave marked B and D are called wave troughs. The maximum
displacement of the string from its equilibrium position denoted yo is called the
amplitude of the wave. The distance between two successive wave crests or
troughs is called wavelength and is denoted with the Greek letter λ.

2
Figure 3: Schematic illustration of a pulse travelling on a string with speed v

Figure 4: Schematic illustration of a wave travelling on a string with speed v


generated by a human hand executing simple harmonic motion

3
One wavelength is sent out by the wave source as its executes one complete
vibration. The time it takes the source to send out one wavelength is called the
periodic time and is denoted τ . During the time it takes the source to send out
one wave length, one wavelength passes through a point P . As a result, point P
undergoes a complete cycle of motion in the same time it takes to complete one
vibration. Thus the period of vibrating source τsource is the same as the period
of the vibration point P in the path of the wave,τwave . This time taken for a
complete vibration of a point P on the wave is called the period of wave and
is also denoted τ .

τsource = τwave (2)


Like for the oscillator, the period of the wave is related to the frequency of
the wave denoted f as
1
f= (3)
τ
The SI units for frequency are Hertz (Hz). Furthermore, the frequency of
the wave can also be defined in terms of the number of wave crests that pass
through a point P each second.
A very important relationship exist between the wavelength and the fre-
quency of a wave. Note that in time period τ the wave travels a distance λ at
a speed v. This distance travelled by the wave λ is related to the periodic time
τ and wave speed v as follows

λ = vτ (4)
1
But recall that τ = , substituting we get
f
1 v
λ=v = (5)
f f
Making the wave speed v the subject of the formula, we get
λ
v = fλ = (6)
τ
This relationship is true for all waves. The frequency of a wave is determined
by the frequency of the source of the wave. The wave speed v is determined by
the properties of the medium through which the wave passes.
The speed of a wave on a string is given by the following relationship without
any derivation. If the tension in a string is T and the mass of length L is m,
then the speed of the wave along the string is given by
s
T
v= (7)
µ
m
where µ = is the mass per unit lengtrh of the string.
L

4
Equating Equation (7) and (7), we get the velocity of a wave travelling on a
string as
s
λ T
v = fλ = = (8)
τ µ
The tension in the string T is responsible for the force that accelerates a piece
of string as the pulse passes through the region. The greater the tension, the
greater the acceleration, and so the motion of the pulse is faster if the tension
is high.
On the other hand, the more massive the string, the more inertia it has. The
mass per unit length µ therefore affects the wave speed v with which a pulse
moves with. A massive string has larger inertia and the speed of the pulse on
it is relatively low.

2 Reflection of Waves
In order for a wave to travel along a string, the string must be held taut at
its right-hand end. A wave travelling to the right cannot continue beyond the
support. So what happens to the kinetic energy and elastic potential energy
carried by the wave since energy cannot be destroyed or just disappear? Two
things may happen:
(i) some of the energy may be absorbed by the material of the support, and
(ii) some of the energy may be reflected back along the string. We assume
that all the energy of the incident waves is reflected and this is a valid
assumption for many cases.
To help us study the reflection of waves, we consider a single pulse propa-
gating along as string as shown in Figure 3(a). Notice that as the pulse moves
along the string, it exerts an upward force on small sections of string where it
has reached displacing that section momentarily from its equilibrium position.
When this pulse reaches the support, it exerts an upward force on the support.
Since the support is fixed in place, it cannot move and therefore does not accel-
erate upwards. However, the support is “unhappy” by the pulse’s attempt to
displace it upwards, therefore according to Newton’s third law of motion, the
support exerts an equal but opposite force downwards on the string. This down-
ward force exerted by the support on the string accelerates the string downwards
to the extent that its linear momentum carries below the equalibrium position.
The result is that the pulse is turned upside down (i.e. inverted) as it hits the
support as shown in Figure 3(b), and the reflected pulse appears as shown in
Figure 3(c). If the string had been completely free to move up and down at
its right end, the pulse would not have been inverted, although it would still
be reflected, since the energy could not just disappear at the end of the string.
In summary, a pulse is inverted by reflection at a fixed end, and it is
reflected but not inverted at the free end.

5
Next, we consider what happens when a reflected pulse travelling to the
left along the string meets an incident pulse moving to the right along the
same string as shown in Figure 5. If the reflected pulse is not inverted as
shown in Figure 5(a), as the incident and reflected pulses begin to overlap, the
displacements of the string at the point of meet is the individual displacement
or amplitude of the pulses are added up which is a vector summation process.
The net displacement is the vector sum of the individual pulse displacements.
Similary, in shown in Figure 5(b), where the reflected pulse is inverted, the
displacement of the inverted pulse is negative but equal in magnitude to that of
the incident pulse. Therefore, as the inverted reflected pulse and the incident
pulse overlap, the net displacement is zero. This kind of interference which
occurs when an incident pulse and a reflected pulse interact is an example of
the principle of superposition which states: a point subjected to two or
more wave pulses simultaneously is displaced an amount equal to the
vector sum of the individual displacements.

Figure 5: Interference of a reflected pulse moving to the left and incident pulse
moving to the right along the same string

The principle of superposition can also be extended and applied to the study
of sinusoidal waves travelling down along string held taut at its right-hand end
and reflected by a rigid support at its end. Recall that sinusoidal waves consist
of a series of pulses propagating along the string, thus whenever an incident
pulse reaches the support, the pulse is inverted and reflected. Next the reflected
pulses now moving to the left begin to interfer with the incident pulses moving
to the right. At the support, the displacement will be zero, as it must always
be.
As the inverted reflected pulses move to the left, there will be several points
along the string where the net or resultant displacement will be zero at any
instant. These points along the string that never move are called nodes and
are denoted N . In between the nodes, different points will move with different
displacements either upwards or downwards. However, exactly midway between

6
the nodes there is one point that always moves the most i.e oscillates the most.
This point is referred to as an antinode and is denoted A. This type of vibra-
tion in which a string vibrates back and forth within a well-defined envelope is
called a standing wave. Standing waves on a string also referred to as wave
resonance occur on a vibrating string when incident and reflected waves pre-
cisely reinforce one another. The combined incident and reflected waves give
rise to nodes and antinodes on the string.

Figure 6: Standing waves occur when a vibrating string with incident and re-
flected waves precisely reinforce one another. The combined incident and re-
flected waves give rise to nodes and antinodes on the string.

If you look at the standing waves shown in Figure 6, you will see that the
λ
nodes are apart. Similarly, the distance between two adjacent antimodes is
2
λ
. Note that the distance between two adjacent nodes or antinodes in
2
λ
a standing wave is always .
2

3 Wave Resonace : Standing Waves


Recall from our study of simple harmonic motion that any vibrating system
has a natural frequency of vibration. If we force such a system to vibrate at its
natural frequency then it undergoes resonance, that is, it vibrates most strongly
when the frequency of the driving force matches the natural frequency of the
system. Similarly, if you vibrate a string with too low frequency, the string
vibrates so little that it appears motionless. If you slowly increase the frequency
of vibration, the string begins to vibrate widely at a certain frequency. This
frequency is called the fundamental resonance frequency and is denoted
f1 . At this fundamental resonance frequency, the string vibrates widely and
appears as a blur. Experiments on vibrations of strings show that a string
can also resonate to other higher frequencies which are integer multiples of the
fundamental resonance frequency such as 2f1 , 3f1 , 4f1 and so on.

7
Figure 7: Schematic illustration of standing waves or wave resonance on a taut
string

During resonance, the string always vibrates in whole segments, where a


segment is defined as the distance between two adjacent nodes or antinodes.
The fixed ends of the string are always nodes. Therefore, the string resonates
only if it is one segment long, two segments long, three segments long and so on.
λ
Since the length of segment is , however, the string can resonate only if its is
   2
λ λ λ
long, 2 long, or 3 long and so on. In general, we can state that a
2 2 2
λ
string fastened firmly at its two ends resonates only if it is a whole number of
  2
λ λ
long. In other words, the string will resonate only if it is long, 2 long,
  2 2
λ
or 3 long and so on. Thus for resonance to occur on a string fastened at
2
both ends, the following condition must be met:
 
λn
L=n , where n = 1, 2, 3, ... (9)
2
Also, recall that the wavelength λ of a wave propagating along the string,
the wave speed v and the frequency of the wave f are related as follows:

v = fλ (10)
When we make the frequency f the subject of the formula, we obtain

8
v
f= (11)
λ
Thus, we see that if a string vibrates with n segments, its wavelength can
be obtained as
2L
λn = , where n = 1, 2, 3, ... (12)
n
For n = 1, we get the wavelength λ1 as
2L
λ1 = = 2L (13)
1
The frequency corresponding to λ1 is
v v
f1 == (14)
λ1 2L
We see that each number of segments n has a corresponding value of fre-
quency of vibration fn .
v v nv  v 
fn = = = =n (15)
λn 2L 2L 2L
n
where
λ1 = 2L is the wavelength
Substituting Equation (14) into Equation (15), we get the allowed frequen-
cies for a string to vibrate with each of the n segments as
fn = nf1 (16)
where
v
f1 = is the fundamental resonance frequency
2L
We notice that the string vibrates at only very special frequencies. In Physics
we say that the frequencies where resonance occurs are quantized meaning that
they occur at discrete values, separated by frequency gaps. These discrete values
of resonance frequency fn are integer multiples of the fundamental resonance
frequency.

fn = nf1 , where n = 1, 2, 3, ... (17)


Often the resonance frequency of a taut string is connected to the music of
string instruments. The fundamental resonance frequency f1 is referred to
as the first harmonic, with f2 as second harmonic, f3 as third harmonic,
f4 as fourth harmonic and fn as nth harmonic. The term harmonic is
used to refer to a vibration of a single sinusoidal wave frequency and the term
simple harmonic motion refers to periodic motion that can be described by
sine and cosine functions of a single frequency.

9
4 Type of Waves
4.1 Transverse and Longitudinal Waves
Waves propagating in a medium can generally be divided into two categories:
transverse and longitudinal waves. Waves of a string are examples of trans-
verse waves. This is because the particles in the medium through which the
wav passes e.g. the string move perpendicular (or tranverse) to the direction of
wave propagation. As the wave on a string propagates from left to right, the
string moves up and down.

Figure 8: Schematic illustration of transverse waves on a string and longitudinal


waves on a spring

Another type of wave called longitudinal waves are produces from during
the force compression of a string tied on the right to the fixed end. If a com-
pressive force is applied on the left side to the spring as shown, the loops near
the end where the compressive force is applied and compressed before the rest
of spring experiences the disturbance or pulse. The compressive loops in turn
exerts a force on the loops to the right of them, and the compression travels
down the spring.
 
λn
L=n , where n = 1, 2, 3, ... (18)
2
When the compression reaches the fixed end at the right, the compression
energy is reflected, thus the compression is reversed and ends up travelling to to
the right. In this type of wave, the particles in the spring move back and forth
in the same direction in which the wave is propagated along the spring. This
kind of compressive waves in which the motion of particles is along the direction
of wave propagation are called longitudinal waves.
When we create a continuous longitudinal wave by connecting the free end of
the spring to a vibrating source that alternatively pushes and pulls the end with

10
frequency f . Regions of closely spaced coils are sent down the spring attenating
with regions of stretched out coils. If the vibrating sources acauses the end
of the spring to move in simple harmonic motion, then we the variation in the
extension (stretch out and compression loops of the coil follow a sinusoidal wave.
This pattern of compression and extension travels along the spring with
velocity v determined by the properties of the spring, the wavelength λ of a
wave i.e. the distance between two successive extension or compressions. The
amplitude is the difference between two adjacent loops or coil distance of
maximum compression (or maximum extension) and the adjacent coil distance
of the undisturbed spring. The same relationship between between the wave
speed v, frequency f and wavelength λ holds:

v = fλ (19)

Sound waves are an important example of longitudinal waves.

4.2 Standing Compressional Waves on a Spring


A longitudinal wave has many features in common with transverse waves on a
string. If a longitudinal wave is sent down a spring, the wave and its energy are
reflected at the end of the spring. This reflected wave can interfere with later
waves being sent down the spring by the source from the source of vibration. If
a proper relation is maintained between the frequency of the source of vibration
and the various parameters of the spring, then resonance occurs.
As with resonance on a string, the source of vibration in the spring system
is usually close to a node since at resonance the spring moves much more than
the source. Also, if the other end of the spring is held motionless, it too must be
at a mode. The displacements on the spring along the x-axis vary sinusoidally
with x. Nodes will occur along the spring at points where waves travelling to
the right and to the left cancel out, leaving the spring neither compressed nor
extended. Standing waves or resonance occurs when the length of the spring is
λ
an integral multiple of the distance between two successive nodes i.e. . The
2
condition for resonance for longitudinal waves on a spring fixed at both ends is
the same as for transverse waves and is given by
 
λn
L=n where n = 1, 2, 3, ... (20)
2
We get the wavelength λ for n segments as
2L
λn = where n = 1, 2, 3, ... (21)
n
Again, recall that the wavelength λ, wave speed v and frequency of any wave
f are related as follows:

v = fλ (22)

11
Making the frequency f the subject of the formula, we get
v
f= (23)
λ
Substituting λn , we get the frequency f as
v
fn =   (24)
2L
n
v
fn = n where n = 1, 2, 3, ... (25)
2L

12

You might also like