PHY1163-Notes-2022 (1)-1
PHY1163-Notes-2022 (1)-1
PHY1163-Notes-2022 (1)-1
Contents I
Physical Constants 1
I MECHANICS 2
1 INTRODUCTION 3
1.1 Physics and Measurement . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Standards and units . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2.1 Fundamental quantities and their measurements . . . . . . . . . . 4
1.2.2 Rules and Conventions in measurement . . . . . . . . . . . . . . . 6
1.2.3 Standards prefixes . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.3 Dimensional Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.4 Uncertainty, Precision and Significant Figures . . . . . . . . . . . . . . . 8
1.5 Order of Magnitude Calculations . . . . . . . . . . . . . . . . . . . . . . 10
1.5.1 Significant Figures and Numbers . . . . . . . . . . . . . . . . . . 11
1.5.2 Rounding Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.6 Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.6.1 Scalar and vector quantities . . . . . . . . . . . . . . . . . . . . . 11
1.6.2 Some Vector Properties . . . . . . . . . . . . . . . . . . . . . . . . 12
1.6.3 Components and magnitude of a vector . . . . . . . . . . . . . . . 12
1.6.4 Vector addition . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.6.5 Unit Vector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.6.6 Vector multiplication . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.6.7 Derivative of a vector . . . . . . . . . . . . . . . . . . . . . . . . . 18
2 KINEMATICS 19
2.1 Motion in One Dimension . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.1.1 Position . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.1.2 Distance and Displacement . . . . . . . . . . . . . . . . . . . . . . 19
2.1.3 Velocity and Acceleration . . . . . . . . . . . . . . . . . . . . . . 20
2.1.4 Acceleration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.1.5 Uniformly Accelerated Motion . . . . . . . . . . . . . . . . . . . . 22
2.1.6 Falling Bodies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.1.7 Variable Acceleration . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.1.8 Exercise . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.2 Motion in Two Dimensions . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.2.1 Displacement, Velocity, and Acceleration Vectors . . . . . . . . . 26
2.2.2 Motion in Two Dimension with Constant Acceleration . . . . . . 28
2.2.3 Projectile Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.2.4 Circular Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.2.5 Relative Velocity and Relative Acceleration . . . . . . . . . . . . 32
I
II Lecture Notes of Physics for Engineers I Module
3 DYNAMICS OF A PARTICLE 34
3.1 Newton’s Laws of Motion . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.1.1 Newton’s First Law of Motion . . . . . . . . . . . . . . . . . . . . 34
3.1.2 Newton’s Second Law of Motion . . . . . . . . . . . . . . . . . . . 34
3.1.3 Newton’s Third Law of Motion . . . . . . . . . . . . . . . . . . . 35
3.1.4 Application of Newton’s laws of motion . . . . . . . . . . . . . . . 35
3.2 Circular Motion and Gravitation . . . . . . . . . . . . . . . . . . . . . . 37
3.2.1 Circular motion dynamics . . . . . . . . . . . . . . . . . . . . . . 37
3.2.2 Newton’s Law of Gravitation . . . . . . . . . . . . . . . . . . . . 37
3.2.3 Satellite Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.3 Work, Energy and Power . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.3.1 Work Done by a Constant Force . . . . . . . . . . . . . . . . . . . 39
3.3.2 Work Done by a Varying Force – One Dimensional Case . . . . . 40
3.3.3 Work Done by a Spring . . . . . . . . . . . . . . . . . . . . . . . 41
3.3.4 Work and Kinetic Energy . . . . . . . . . . . . . . . . . . . . . . 41
3.3.5 Power . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.4 Potential Energy and Conservation of Energy . . . . . . . . . . . . . . . 44
3.4.1 Conservative and Nonconservative Forces . . . . . . . . . . . . . . 44
3.4.2 Potential Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.4.3 Conservation of Mechanical Energy . . . . . . . . . . . . . . . . . 47
3.4.4 Gravitational Potential Energy near the Earth’s Surface . . . . . 47
3.4.5 Nonconservative Forces and the Work-Energy Theorem . . . . . . 49
3.4.6 Potential Energy Stored in a Spring . . . . . . . . . . . . . . . . . 49
3.4.7 Relationship between Conservative Forces and Potential Energy . 51
3.5 Linear Momentum and Collisions . . . . . . . . . . . . . . . . . . . . . . 51
3.5.1 Linear Momentum and Impulse . . . . . . . . . . . . . . . . . . . 51
3.5.2 Conservation of Linear Momentum for a Two-Particles System . . 53
3.5.3 Concept of Collision . . . . . . . . . . . . . . . . . . . . . . . . . 55
3.5.4 Collisions in One-Dimensional Space . . . . . . . . . . . . . . . . 56
3.5.5 Collisions in Two-Dimensional Space . . . . . . . . . . . . . . . . 58
6 FLUID MECHANICS 83
6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
6.1.1 States of matter . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
6.1.2 Properties of Fluids . . . . . . . . . . . . . . . . . . . . . . . . . . 84
6.2 Pressure in Fluids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
6.2.1 Definition of the Pressure . . . . . . . . . . . . . . . . . . . . . . 86
6.2.2 Variation of pressure with depth: Pascal’s law . . . . . . . . . . . 87
6.2.3 Pressure measurements . . . . . . . . . . . . . . . . . . . . . . . . 88
6.2.4 Buoyant forces and Archimedes’ principle . . . . . . . . . . . . . . 88
6.3 Fluid dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
6.3.1 Flow Characteristics . . . . . . . . . . . . . . . . . . . . . . . . . 90
6.3.2 Streamlines and equation of continuity . . . . . . . . . . . . . . . 91
6.3.3 Bernoulli’s equation . . . . . . . . . . . . . . . . . . . . . . . . . . 91
6.3.4 Poiseuille’s Law and Stokes’ Law . . . . . . . . . . . . . . . . . . 92
Bibliography 121
Physical Constants
1
Part I
MECHANICS
2
Chapter 1
INTRODUCTION
3
4 Lecture Notes of Physics for Engineers I Module
Physical quantities are classified into two categories: fundamental quantities and
delivered quantities.
Fundamental quantities are quantities which cannot be expressed in terms of any physical
quantity. In other hand, derived quantities are quantities that can be expressed in terms
of fundamental quantities. In mechanics, the three fundamental quantities are length,
mass, and time. All other quantities in mechanics can be expressed in terms of these
three.
If we are to report the results of a measurement to someone who wishes to reproduce this
measurement, a standard must be defined.
The measurement of quantity is done by comparing it with some standard called unit. A
unit, therefore, is any division which is accepted as one unit of that quantity. For example,
if someone familiar with our system of measurement reports that a wall is 2 meters high
and our unit of length is defined to be 1 meter, we know that the height of the wall is
twice our basic length unit.
Whatever is chosen as a standard must be readily accessible (permanence or accessibility)
and must possess some property that can be measured reliably. Measurement standards
used by different people in different places throughout the Universe must yield the same
result (reproducibility). In addition, standards used for measurements must not change
with time (invariability). The System International (SI) of units is logically far superior
to all other system.
• The time which the unit is second (s). The second used to be defined by Earth’s
rotation, but that’s not constant, so it was redefined as 1/86400 of the mean solar day
in year 1900. In 1967 the second was given an new operational definition involving
atomic vibrations. That was defining the second as the duration of 9,192,631,770
periods of the radiation corresponding to the transition between the two hyperfine
levels of the ground–state of the caesium-133 atom. But this definition was reviewed
a bit in 2018 when the CGPM decided to base the definition of SI on fixed 7 defining
constants. So the second is defined by taking the fixed numerical value of the caesium
frequency ∆vCs , the unperturbed ground–state hyperfine transition frequency of the
caesium 133 atom, to be 9 192 631 770 when expressed in the Hz, which is equal to
s−1 . The device that implements this definition—which will seem a lot less obscure
once you’ve studied some atomic physics—is called an atomic clock. Thus, the
second is SI unit of the time, the measure of the duration between two events.
• The length which the unit is meter (m). The meter was first defined as one
ten-millionth of the distance from the equator to the north pole. In 1889 a standard
meter was fabricated to replace the Earth-based unit, and in 1960 that gave way
to a standard based on the wavelength of light. Such an operational definition, a
measurement standard based on a laboratory procedure, has the advantage that
scientists anywhere can reproduce the standard meter. By the 1970s, the speed of
light had become one of the most precisely determined quantities. As a result, in
1983 the meter was given a new operational definition, defining the meter as the
length of the path traveled by light in vacuum during a time interval of 1/299,792,458
of a second. This definition was reviewed in 2018. Now the meter is defined by
taking the fixed numerical value of the speed of light in vacuum c to be 299 792 458
when expressed in the unit m s1 , where the second is defined in terms of ∆vCs . Thus,
the meter is the SI unit of the length. The length can be defined as the distance
between two points in space.
• The mass which the unit is kilogram (kg). Until November 2018, the mass standard
was the least satisfactory. Unlike the operational definitions of length and time,
based on procedures that can be repeated anywhere, the unit of mass was defined by
a particular object—the international prototype kilogram kept at the International
Bureau of Weights and Measures at Sèvres, France. The prototype kilogram was
made of a special platinum-iridium alloy that is very hard and not subject to
corrosion. Nevertheless, it could change, and in any event comparison with such a
standard was less convenient than an operational definition that can be checked in
a laboratory. So scientists developed techniques based on counting the atoms in a
given volume, to scale up from the mass of a single atom to a new definition of the
kilogram. Scientists adopted a new definition of the kilogram in 26th CGPM (2018).
Since then, the kilogram is defined by taking the fixed numerical value of the Planck
constant h to be 6.626 070 15 ×10−34 when expressed in the unit J s, which is equal
to kg m2 s−1 . Thus, the kilogram is the SI unit of the mass. The mass of a body is
defined as the quantity of matter in the body, and can never be zero.
• The electric current which the unit is ampere (A). The ampere, symbol A, is
defined by taking the fixed numerical value of the elementary charge e to be 1.602
176 634 ×10−34 when expressed in the unit C, which is equal to A s, where the
second is defined in terms of ∆vCs .
5
6 Lecture Notes of Physics for Engineers I Module
The last comment leads naturally to the subject of dimensional analysis. i.e, the use
of the idea of dimensional consistency to guess the form of simple laws of
physics.
Example:
The speed of sound in a gas might plausible depend on the pressure p, the density ρ, and
the volume V of the gas. Use dimensional analysis to determine the exponents x, y and z
in the formula v = Cpx ρy V z ; where C is a dimensionless constant.
7
8 Lecture Notes of Physics for Engineers I Module
Answer:
Equating the dimension of both sides of the above equation, we obtain
L M x M y 3 z
T
= T 2L L3
(L )
A comparison of the exponents of L, M and T on either side of the above expression
yields
1 = −x − 3y + 3z
x = 1/2
0=x+y ⇔ y = −1/2
−1 = −2x z=0
q
p
Hence, v = C ρ
Measurements always have uncertainty mainly due to the quality of the apparatus, the
skills of the experimenter and the number of measurement performed.
Suppose someone is measuring the area of a rectangular plate using a meter stick, with
±0.1 cm as accuracy, also called ”absolute uncertainty”.
Theorems of Uncertainties
∆G
∆Grel =
G
In the above example, the relative uncertainty of the length is
∆L 0.1 cm
∆Lrel = = = 0.00613 ≈ 0.61%of uncertainty or 99.39% of certitude
L 16.3 cm
In case of measurements calculated using mathematical equations, the theorems of uncer-
tainties are helpful.
Gmax − Gmin X + ∆X + Y + ∆Y − X − ∆X + Y − ∆Y
∆G = ∆G = =
2 2
⇒ ∆G = ∆X + ∆Y
If we have G = X − Y , with X = X ± ∆X and Y = Y ± ∆Y , the absolute uncertainty
∆G is, by definition:
Gmax − Gmin X + ∆X − Y − ∆Y − X − ∆X − Y + ∆Y
∆G = ∆G = =
2 2
⇒ ∆G = ∆X + ∆Y
Conclusion: The absolute uncertainty on an addition and on a subtraction is the same
and it is equal to the sum of the absolute uncertainties on the terms which are added or
subtracted.
The relative uncertainty on the sum and difference become, by definition:
∆G ∆X + ∆Y ∆G ∆X + ∆Y
∆Grel = = and ∆Grel =
G X +Y G X −Y
respectively.
∆G X · ∆Y + Y · ∆X ∆X ∆Y
∆Grel = = = +
G X ·Y X Y
X
If we have G = Y
; with X = X ± ∆X and Y = Y ± ∆Y , the absolute uncertainty ∆G is,
by definition:
Gmax − Gmin 1 X + ∆X X − ∆X
∆G = ∆G = = −
2 2 Y − ∆Y Y + ∆Y
" #
1 (X · Y + Y · ∆X + X · ∆Y + ∆X · ∆Y ) − (X · Y − Y · ∆X − X · ∆Y + ∆X · ∆Y )
= 2
2 Y − (∆Y )2
9
10 Lecture Notes of Physics for Engineers I Module
X · ∆Y + Y · ∆X
∆G = 2
Y
∆G X · ∆Y + Y · ∆X Y ∆X ∆Y
∆Grel = = 2 · = +
G Y X X Y
The determination of uncertainties on integer powers or nth roots is deduced from the
equation for the products:
G = X n ⇒ G = X · X · X · X · · · X (n factors) and
➢ ∆Grel = ∆G
G
= ∆X
X
+ ∆X
X
+ · · · + ∆X
X
= n ∆X
X
√
➢ G = n X = X 1/n ⇒ ∆Grel = ∆G G
= n1 ∆X
X
√
➢ G = n X m = X m/n ⇒ ∆Grel = ∆G G
=m ∆X
n X
➢ In general for polynomial and rational functions, we use the partial derivatives:
G = f (x, y, x, · · · )
∂f ∂f ∂f
⇔ dG = dx + dy + dz + · · ·
∂x ∂y ∂z
and then:
∂f ∂f ∂f
∆G = | | · ∆x + | | · ∆y + | | · ∆z + · · ·
∂x ∂y ∂z
Example: (1) How many gallons do you estimate that a cylindrical container 36 cm in
diameter and 68 cm high will hold?
Solution:
Volume of that cylinder:
V = π × r2 × h = π × (18 cm)2 × 68 cm = 3 × (20 cm)2 × 70 cm = 80, 000cm3 ≈ 80 l
Since a litter is approximately equal to a quater of gallon, it follows that the container
will hold approximately 20 gallons. (note that the exact calculation gives 18.3 gallons)
193.7×39.64 (20×102 )(4×101 )
(2) Computer: 8.71
≈ 9
≈ 1 × 103 (exact calculation is about 882)
1.6 Vectors
1.6.1 Scalar and vector quantities
Some physical quantities are entirely defined by a numerical value and are called scalar
quantities or scalars. Examples of scalars include time, mass (e.g 5 kilograms), tempera-
ture (e.g 22 degrees Celsius), energy and volume. Other physical quantities are defined by
11
12 Lecture Notes of Physics for Engineers I Module
both a numerical value and a direction in space and these are called vector quantities
or vectors. Examples of vectors include force, velocity (e.g. 12 meters per second north),
moment and displacement (e.g. 120 cm at 300 ). The only basic SI unit for vectors is the
meter, so every vector quantity must involve meters in its definition and unit.
• Two vectors that have the same direction are said to be parallel.
• Two vectors that have the same length and the same direction are said to be equal
no matter where they are located.
• The negative of a vector is a vector with the same magnitude but opposite direction.
The magnitude of a vector is a positive number (with units!) that describe its size or its
length.
The coordinate a is the scalar horizontal component of the vector, and the coordinate b
is the scalar vertical component of the vector. By scalar, we mean a numerical quantity
rather than a vector quantity. Thus, is considered to be the component form of ⃗v . Note
that a and b are NOT vectors and should not be confused with the vector component
definition.
Thus, if we agree that the vector is to start at the origin,the positive end may be the
⃗
specified by giving the Cartesian coordinates (Ax , Ay , Az ) of the arrow head. Although A
could have represented any vector quantity (momentum, electric field, velocity, etc.), one
particularly important vector quantity, that is, the displacement from the origin to the
point (x, y, z), is denoted by the special symbol r or ⃗r.
cosα = x/r, cosβ = y/r, and cosγ = z/r are called the direction cosines, with α being
the angle between the given vector and the positive x-axis, and so on. One further bit of
vocabulary: The quantities Ax ; Ay , and Az are known as the (Cartesian) components
of A or the projections of A, i.e.Ax = x; Ay = y, and Az = z
−→
Now consider a vector AC = ⃗v which is not starting from the origin. Let A(x1 , y1 , z1 ) and
−→
C(x2 , y2 , z2 ) be the starting point and the end, respectively, of the vector AC = ⃗v . The
−→
components form of AC are v⃗x , v⃗y and v⃗z . ⃗v =< x2 − x1 , y2 − y1 , z2 − z1 >= v⃗x + v⃗y + v⃗z .
−→
The length, or magnitude, of a vector AC is given by
−→ p q
∥AC∥ = v = (x2 − x1 )2 + (y2 − y1 )2 + (z2 − z1 )2 = vx2 + vy2 + vz2 (1.1)
A unit vector ⃗e can be constructed along a vector using the direction cosines as its
components along the x, y and z directions. For example, the unit-vector ⃗e along the
⃗ is obtained from
vector A
⃗
A A⃗x A⃗y A⃗z
⃗e = = + + (1.2)
A A A A
Tip-to-Tail Method
We can add any two vectors, A and B, by placing the tail of B so that it meets the tip of
A. The sum, A+B, is the vector from the tail of A to the tip of B. Note that you will
get the same vector if you place the tip of B against the tail of A. In other words, A+B
and B+A are equivalent.
Parallelogram Method
To add A and B using the parallelogram method, place the tail of B so that it meets
the tail of A. Take these two vectors to be the first two adjacent sides of a parallelogram,
and draw in the remaining two sides. The vector sum, A+B, extends from the tails of A
and B across the diagonal to the opposite corner of the parallelogram. If the vectors are
perpendicular and unequal in magnitude, the parallelogram will be a rectangle. If the
vectors are perpendicular and equal in magnitude, the parallelogram will be a square.
13
14 Lecture Notes of Physics for Engineers I Module
Of course, knowing what the sum of two vectors looks like is often not enough. Sometimes
you will need to know the magnitude of the resultant vector. This depends not only on
the magnitude of the two vectors you are adding, but also on the angle between the two
vectors. If vectors A and B are not perpendicular, you will calculate the magnitude of
A+B by using the angle α between A and B.
√
A+B = A2 + B 2 + 2AB cosα (1.3)
The second method, you add their components on x and y axis, and you use the
Pythagorean theorem since the projections are perpendiculars.
If the vectors you want to add are in the same direction, they can be added using simple
arithmetic. For example, if you get in your car and drive eight km east, stop for a break,
and then drive six km east, you will be 8 + 6 = 14 km east of your origin. If you drive
eight km east and then six km west, you will end up 8 − 6 = 2 km east of your origin.
• Vector Subtraction
You probably know that subtraction is the same thing as adding a negative: 8 − 5 is the
same thing as 8 + (−5). The easiest way to think about vector subtraction is in terms of
adding a negative vector. What is a negative vector? It is the same vector as its positive
counterpart, only pointing in the opposite direction. Then A-B, is the same thing as
A+(-B). For instance, let’s take the two vectors A and B: To subtract B from A, take a
vector of the same magnitude as B, but pointing in the opposite direction, and add that
vector to A, using either the tip-to-tail method or the parallelogram method.
⃗ = Ax î + Ay ĵ
A (1.4)
Now let us see how to use components to add vectors when the graphical method is not
⃗ to vector A
sufficiently accurate. Suppose we wish to add vector B ⃗ in Equation 1.4, where
vector B⃗ has components Bx and By . The resultant vector R ⃗ =A ⃗+B ⃗ is
Rx = Ax + Bx and Ry = Ay + By (1.6)
Dot product
The dot product, also called the scalar product, takes two vectors, multiplies them
together, and produces a scalar. The smaller the angle between the two vectors, the
greater their dot product will be. A common example of the dot product in action is
the formula for work. Work is a scalar quantity, but it is measured by the magnitude of
force and displacement, both vector quantities, and the degree to which the force and
displacement are parallel to one another.
The dot product is denoted by “•” between two vectors. The dot product of vectors A
and B results in a scalar given by the relation:
15
16 Lecture Notes of Physics for Engineers I Module
⃗•B
A ⃗ = AB cosθ. (1.7)
where θ is the angle between the two vectors. Order is not important in the dot product as
can be seen by the dot products definition. The dot product has the following properties:
⃗a • ⃗a = a2
When working with vectors represented in a rectangular coordinate system by the com-
ponents A⃗ = Ax i + Ay i + Az k and B
⃗ = Bx i + By j + Bz k, then the dot product can be
evaluated from the relation:
⃗•B
A ⃗ = Ax Bx + Ay By + Az Bz (1.8)
This can be verified by direct multiplication of the vectors and noting that due to the
orthogonality of the base vectors of a rectangular system one has i • j = 0 ; j • k =
0 ; i • k = 0 and i • i = 1 ; j • j = 1 ; k • k = 1.
Order is important in the cross product. If the order of operations changes in a cross
product the direction of the resulting vector is reversed. That is,
⃗×B
A ⃗ = −B
⃗ ×A
⃗ = C.
⃗
The cross product of vectors A and B gives a new vector C, perpendicular to both vectors
A and B, and whose magnitude:
⃗ × B∥
∥A ⃗ = ∥C∥
⃗ = AB sinθ (1.9)
⃗ is the area generated by
where θ is the angle between the two vectors. The value of ∥C∥
those vectors as shown on Figure 1.9. The cross product has the following properties:
⃗a × (⃗b + ⃗c) = ⃗a × ⃗b + ⃗a × ⃗c
(⃗a + ⃗b) × ⃗c = ⃗a × ⃗c + ⃗b × ⃗c
When working in rectangular coordinate systems, the cross product of vectors ⃗a and ⃗b
given by ⃗a = ax i + ay j + az k and ⃗b = bx i + by j + bz k, can be evaluated using the rule:
⃗i ⃗j ⃗k
⃗a × ⃗b = ⃗c = ax ay az = (ay bz − by az )⃗i − (ax bz − bx − az )⃗j + (ax by − bx ay )⃗k (1.10)
bx by bz
One can also use direct multiplication of the base vectors using the relations:
i × j = k ; j × k = i ; k × i = j and i × i = 0 : j × j = 0 : k × k = 0
ax ay az
⃗
⃗a •(b×⃗c) = α = bx by bz = (by cz −cy bz )ax −(bx cz −cx −bz )ay +(bx cy −cx by )az . (1.12)
cx cy cz
Verify the expansion of the triple vector product, by direct expansion in Cartesian
coordinates.
(i) ⃗a × (⃗b × ⃗c) = ⃗b(⃗a • ⃗c) − ⃗c(⃗a • ⃗b)
17
18 Lecture Notes of Physics for Engineers I Module
⃗
dA dAx⃗ dAy ⃗ dAz ⃗
= i+ j+ k
du du du du
The derivative of a vector is a vector whose Cartesian components are ordinary derivatives.
It follows that the derivative of the sum of the two vectors is equal to the sum of the
namely vector:
d(A +⃗ B) dA⃗ dB ⃗
= +
du du du
The rules for differentiating vector products obey similar rules of a vector calculus. For
example:
⃗
d(nA) dn ⃗ dA⃗
= A+n
du du du
d(A ⃗• B) ⃗
dA ⃗
⃗ • dA
⃗ +A
= •B
du du du
⃗ B)
d(A × ⃗
dA ⃗
= ×B ⃗ +A ⃗ × dB
du du du
Notice that it is necessary to preserve the order of the terms in the derivative of cross
product.
Kinematics is the part of Classical Mechanics which describes the motion of points, bodies
(objects) and systems of bodies (groups of objects) without consideration of the causes of
that motion.
images/Displacementone.png
2.1.1 Position
Ideally, a particle (mass point) is a body whose dimension can be neglected in a given
situation. The position of a particle is defined related to a reference point called origin.
In a given reference system, the position of a particle can be specified by a single vector,
called vector position of the particle relative to the origin of the coordinate system as seen
in Figure 1.2.
In order to describe the motion of an object, you must first be able to describe its position
- where it is at any particular time. More precisely, you need to specify its position relative
to a convenient reference frame. Earth is often used as a reference frame, and we often
describe the position of an object as it relates to stationary objects in that reference
frame. For example, a rocket launch would be described in terms of the position of the
rocket with respect to the Earth as a whole, while a teacher’s position could be described
in terms of where he is in relation to the nearby (white or black or green) board. In
other cases, we use reference frames that are not stationary but are in motion relative to
the Earth. To describe the position of a person in an airplane, for example, we use the
airplane, not the Earth, as the reference frame.
19
20 Lecture Notes of Physics for Engineers I Module
d
v̄ = (2.1)
t
• Example
How far can a cyclist travel in 4.0 h if his average speed is 11.5 km/h?
Solution:
Since we want to find the distance traveled, we rewrite the equation above. Since
v = 11.5 km/h and t = 4.0h, d=(11.5 km/h)(4.0 h)=46 km.
Velocity
The word Velocity is used to signify both the magnitude (numerical value) of how fast
an object is moving and its direction (it is therefore called a vector ). The velocity of a
body is a vector quantity that describes both how fast it is moving and the direction in
which it is headed.
In the case of a body traveling in a straight line, its velocity is simply the rate at which it
covers distance. The average velocity v̄ is defined as the displacement divided by the
clasped time (which is t = tf − ti )
⃗
D x⃗2 − x⃗1 ∆⃗x
⃗v̄ = = = (2.2)
t t2 − t1 ∆t
• Example
The position of a bowling ball as a function of time is plotted as moving along the
axis of a graph. At ti = 3.00 s, its position is xi = 40.5 m; at tf = 5.50 s, its
position is xf = 18.2 m. What was its average velocity?
Solution:
∆x x2 −x1
v̄ = ∆t
= t2 −t1
;
∆x = x2 − x1 = 18.2 m − 40.5 m = −22.3 m and ∆t = t2 − t1 = 5.5 − 3 = 2.5 s
∆x −22.3 m
v̄ = ∆t
= 2.5 s
= −8.92 m/s
The displacement and the velocity are negative, so the ball is moving to the left
along the axis.
Instantaneous velocity
The instantaneous velocity at any point is the average velocity over an indefinitely short
time interval;
∆⃗x d⃗x
⃗v = lim = (2.3)
∆t→0 ∆t dt
This Equation (2.3) is the definition of instantaneous velocity for one-dimensional motion.
In general,
Figure 2.3 shows that as ∆t approaches zero, the point P ′ approaches the point P , and
the direction of the velocity d⃗
r
dt
approaches the direction of the tangent to the path P .
Therefore. the velocity vector, ⃗v is always tangent to the path of the motion.
d⃗r dx dy dz
⃗v = = ⃗i + ⃗j + ⃗k = vx⃗i + vy⃗j + vz⃗k (2.4)
dt dt dt dt
The magnitude of the velocity is called the speed. In rectangular components, the speed is
q
v = ∥⃗v ∥ = vx2 + vy2 + vz2 (2.5)
• Example
A particle is moving along the x axis. Its position as a function of time is given by
the equation x = At2 + B where A = 2.10 m/s2 and B = 2.80 m.
(a) Determine the displacement of the particle during the time interval from ti =
3.00 s to tf = 5.00 s.
(b) Determine the average velocity during this time interval.
(c) Determine the magnitude of instantaneous velocity at t = 5.00 s.
Solution:
(a) At ti = 3.00 s, the position is xi = At2 + B = (2.10 m/s2 )(3 s)2 + 2.8 m = 21.7 m
At tf = 5.00 s, the position is xt = At2 + B = (2.10 m/s2 )(5 s)2 + 2.8 m = 55.3 m
The displacement is thus xf − xi = 55.3 m − 21.80 m = 33.6 m
xt −xi
(b) The magnitude of the average velocity is v = tf −xi
= 33.6/2.00 = 16.8 m/s
(c) Let us determine the instantaneous velocity for any time t (that is, determine v as
2
a function of t). v = dr
dt
= d(Atdt+A) = 2At = (4.2 m/s2 )t, where we have substituted
in A = 2.10 m/s2 . At t = 5.0 s, we have v = (4.20 m/s2 )(5.00 s) = 21.0 m/s.
21
22 Lecture Notes of Physics for Engineers I Module
2.1.4 Acceleration
An object whose velocity is changing in time is said to be accelerating. A car whose
velocity increases from zero to 80 km/h is accelerating, if a second car accomplishes this
change in velocity in less time than the first car, it is said to undergo a great acceleration.
In general, the average acceleration, ā, during a time interval ∆t = tf − ti during which
the velocity changes by ∆v = vf − vi is defined as:
∆⃗v d⃗v
⃗a = lim = (2.7)
∆t→0 ∆t dt
Figure 2.4: Graph representation of (a) The Velocity versus the Time; (b) The Acceleration
versus the Time
A common problem is to determine the velocity of an object after a certain time given its
acceleration. We can solve such problem by solving for v in the last equation:
vf = vi + a(tf − ti ) (2.8)
For example, it may be known that the acceleration of a particular motor-cycle is 4.0m /s2
and we wish to determine how fast it will be going after, 6.0 s. Assuming it starts from rest
(vi = o), after 6 s the velocity will be v = a(tf − ti ) = (4.0 m/s2 )(6.0 s − 0 s) = 24 m/s.
Next, let us see how to calculate the position of an object after a time t when it is
x −x
undergoing constant acceleration. From the definition of average velocity, v̄ = tff −tii ,
xf = xi + v̄(tf − ti ) (2.9)
Because the velocity increases at uniform rate, the average velocity, v̄, will be midway
between the initial and final velocity,
vf + vi
v̄ = (2.10)
2
Combining the last three equations (2.8 - 2.10), we find
vf + vi (vi + a(tf − ti )) + vi
xf = xi + ( )(tf − ti ) = xi + ( )(tf − ti )
2 2
1
xf = xi + (vf − vi )(tf − ti ) + a(tf − ti )2 (2.11)
2
We now derive the fourth equation, which is useful in a situation when, for instance, the
acceleration, position, and initial velocity are known, and the final velocity is desired but
the final time is not known. To obtain the velocity at any time in term of vi , a, xi and xf .
From Equation (2.8), we get tf − ti = (vf − vi )/a, and substituting this into the above
Equation (2.9) and substituting v̄ by Equation (2.10) we get:
vf + vi
xf = xi + (vf − vi )/a
2
Solution:
We use vf2 = vi2 + 2a(xf − xi ), with xi = 0, vi = 0, xf = 100 m, and a = 12.0 m/s2 ,
then
√
vf2 = 0 + 2(12.0 m/s2 )(100 m) = 2400 m2 /s2 =⇒ vf = 2400 = 49 m/s.
Unfortunately, this length runway is not sufficient. By solving vf2 − vi2 = 2a(xf − xi )
you can determine how long a runway is needed for this plane.
23
24 Lecture Notes of Physics for Engineers I Module
We call this acceleration the acceleration due to gravity and we give it the symbol g.
Its value is approximately g = 9.80 m/s2 , and varies slightly according to latitude and
altitude but these variations are so small that are ignored, the effect of air resistance are
often small so that are neglected. When a falling body has the constant acceleration g
and the formulas for uniformly accelerated motion apply. Thus a body dropped from rest,
after the time t has the velocity
v = vo + gt (2.13)
1
v = vo − gt and h = vo t − gt2
2
With these formula, you can find the time (ts )taken to reach its highest point,
vo
ts = , (2.15)
g
vo
ts = 2 . (2.17)
g
• Example
A student throws a ball up in the air with an initial velocity of 12 m/s and then
catches it as it comes back down to him. What is the ball’s velocity when he catches
it? How high does the ball travel? How long does it take the ball to reach its highest
point?
Figure 2.5: (a)Three graphical views of uniformly accelerated motion; (b) Freely falling
object
• Example
An object starts from rest (v1 = 0) at t1 = 0 and accelerates at a rate given by
a(t) = (7.00 m/s3 )t. What is (a) its velocity and (b) its displacement 2.00 s later?
Solution
(a) Let us first determine v as a function of t. We set ti = 0; tf = t; vi = 0, vf = v(t).
Z tf Z t
t2 t2
v(t) = a(t) dt = 7t dt = 7 |t0 = 7( − 0) = (3.5 m/s3 )t2
ti 0 2 2
At t = 2.00 s, v = (3.50 m/s3 )(2.00 s)2 = 14.00 m/s
(b) To get the displacement we use:
Z tf
x(t) = v(t) dt,
ti
2.1.8 Exercise
1. A stone is thrown from the top of a building with an initial velocity of 20 m/s
straight upward. The building is 50 m high, and the stone just misses the edge of
the roof on its way down. Taking the origin pf axis of the ground, determine: (a)
The time t1 needed for the stone to reach its maximum height; (b) the maximum
height Hm from the ground; (c) the time to needed for the stone to return to the
level of the thrower; (d) the velocity vo of the stone at this instant; (e) the velocity
v and position y of the stone at t = 5 s; (f) the velocity v2 of the stone just before it
hits the ground; and (g) the time t2 the stone is in the air.
2. A particle moves along the positive x axis in such a way that its coordinate varies
in time according to the expression x = 4 + 2t − 3t2 , where x is in m and t is in s.
(a) Make a graph of x versus t for the interval t = 0 to t = 2 s; (b) Determine the
initial position and initial velocity of the particle; (c) Determine at what time the
particle reaches a maximum position coordinate (Note that at this time v = 0); (d)
Calculate the coordinate, velocity, and acceleration at t = 2 s.
25
26 Lecture Notes of Physics for Engineers I Module
images/two_dim.png
As the particle moves from P to Q in the time interval ∆t = tf − ti , the position vector
changes from ⃗ri to ⃗rf where the indices i and f refer to the initial and final values,
respectively. Because ⃗rf = ⃗ri + ∆⃗r, the displacement vector for the particle is given by
We now define the average velocity vector of the particle during the time interval ∆t
as the ration of the displacement to the time interval if this displacement:
⃗v = ∆⃗r (2.19)
∆t
Department of Physics, 2019-2020
Section 2.2: Motion in Two Dimensions 27
Since the displacement is a vector and the time interval is a scalar, we conclude that the
average velocity vector is a vector quantity directed along ∆⃗r.
The instantaneous velocity vector, ⃗v , is defined as the limit of the average velocity
vector, ∆⃗
r
∆t
, and ∆t approaches zero:
∆⃗r d⃗r
⃗v = lim = (2.20)
∆t→0 ∆t dt
The direction of the velocity vector is along a line that is tangent to the path of the
particle and in the direction of motion.
The average acceleration vector of the particle as it moves from P to Q is defined as
the ration of the change in the instantaneous velocity vector, ∆⃗v = ⃗vf − ⃗vi , to the elapsed
time, ∆t = tf − ti as shown in Figure 2.7
images/chnge_in_velocity.png
Since the average acceleration is the ratio of a vector , ∆⃗v , and a scalar, ∆t, we conclude
that ⃗a is a vector quantity directed along ∆⃗v .
The instantaneous acceleration vector, ⃗a, is defined as the limiting values of the
ratio ∆⃗
v
∆t
as ∆t approaches zero:
∆→−
v d⃗v d2⃗r
⃗a = lim = = 2 (2.22)
∆t→0 ∆t dt dt
In other words, the instantaneous acceleration vector equals the first derivative of the
velocity vector with respect to time.
27
28 Lecture Notes of Physics for Engineers I Module
d⃗r dx dy
⃗v = = ⃗i + ⃗j ⇒ ⃗v = vx⃗i + vy⃗j (2.24)
dt dt dt
Because ⃗a is constant, its components ax and ay are also constants. Therefore, we can
apply the equations of kinematics to both the x and y components of velocity vector.
Substituting vx = vxo + ax t and vy = vyo + ay t into Equation 2.24 gives
⃗v = vxo⃗i + vyo⃗j + ax⃗i + ay⃗j t ⇒ ⃗v = ⃗vo + ⃗at (2.25)
Similarly, from kinematics we know that the x and y coordinates of a particles moving
with constant acceleration are given by
(
x = xo + vxo t + 12 ax t2
y = yo + vyo t + 12 ay t2
⃗ ⃗
⃗ ⃗
1 ⃗ ⃗
1
⃗r = xo i + yo j + vxo i + vyo j t + ax i + ay j t2 ⇒ ⃗r = ⃗ro + ⃗vo t + ⃗at2 (2.26)
2 2
This equation says that the displacement vector ⃗r − ⃗ro is the vector sum od a displacement
vector ⃗vo t, arising from the initial velocity of the particle, and a displacement vector 12 ⃗at2 ,
resulting from the uniform acceleration of the particle.
Since the equations 2.25 and 2.26 are vector expressions having one or more components,
we may write the component forms of these expressions along the x and y axes with
⃗ro = ⃗0:
(
vx = vxo + ax t
⃗v = ⃗vo + ⃗at ⇒
vy = vyo + ay t
and
(
1 x = vxo t + 21 ax t2
⃗r = ⃗vo t + ⃗at2 ⇒
2 y = vyo t + 12 ay t2
We will describe the motion in the Cartesian coordinate two dimensional system x, y,
with y axis being oriented vertically. The small object was thrown from the position
x = 0, y = 0 of this coordinate system with the initial velocity, ⃗v0 at an angle α from the
positive direction of x axis. The projectile will move along the trajectory drawn by the
dotted line in Figure 2.8. This trajectory is a part of the parabola and this will be evident
from equations we will derive here.
To study the motion of this object we decompose the initial velocity ⃗v0 into its Cartesian
components vx0 and vy0 . These components are also vectors, but we omit arrows in this
notation because subscript x and y implies values with direction. From the trigonometry
it follows that:
v0x = v0 cosα and v0y = v0 sinα (2.27)
The motion along the x axis is completely independent from the motion along the y axis.
Along the horizontal x axis there is a motion with constant velocity v0x . The distance
traveled along the x axis as a function of time can be calculated using:
There are two main questions asked for such a case of projectile motion:
(1) What is a largest vertical displacement (Maximum height of a projectile) H for
a given angle α ?
(2) What is a largest horizontal displacement (Horizontal range for a projectile) R
for a given angle α?
The rising of the projectile continues till the moment its vertical component of velocity
decreases to zero, so we have 0 = v0y − gtr so tr = v0y /g, i.e. the time of rising of the
projectile tr is given by
29
30 Lecture Notes of Physics for Engineers I Module
v0 sinα
tr = (2.31)
g
Substituting this Equation (2.31) for rising time into Equation (2.29) which describes the
vertical displacement, we get after very little algebra, the largest vertical displacement
obtained by the projectile
1 2
H = ymax = v sin2 α (2.32)
2g 0
We already calculated the time of rising to the height H. It is easy to that the time of
falling tf from this height is equal to the time of rising, that is tf = tr . So the total time
(the time of flight) tT the projectile traveling in the x direction is
vo sinα
tT = 2tr = 2tf = 2 (2.33)
g
Substituting this expression for time of motion equation for the largest horizontal dis-
placement, R, we obtain R = v0 cosα(2vo sinα)/g. From Equation (??), we know that
2sinαcosα = sin(2α); thus
v 2 sin(2α)
R= 0 (2.34)
g
The next question we can ask is: for which angle α, with a given initial velocity v0 , will
Figure 2.9: The maximum horizontal displacement. A projectile launched from the origin
with an initial speed of 50 m/s at various angles of projection. [?]
The first solution, α = 00 leads to vertical displacement equal to zero – the minimum of
vertical displacement. Therefore the maximum vertical displacement is obtained for the
firing angle α = 900 .
The angle at which the maximum horizontal displacement is obtained can be found if
sin(2α) = 1, i.e. 2α = 900 , so so, α = 450 . Note that complementary values of angles
result in the same value of R (range of the projectile). If we express vertical displacement
y as a function of horizontal displacement x, we obtain:
g
y(x) = tanα x − x2 (2.35)
2 v02 cos2 α
• Example:
A football is thrown with a velocity of 10 m/s at an angle of 300 above the horizontal.
(a) How far away should its intended receiver be? (b) What will the time of flight
be? (c) What is the velocity (magnitude and direction) after 0.6 s?
Figure 2.10: Velocity and acceleration vectors for an object in circular motion
2.2.4 Circular Motion
Circular motion is a motion of an object along the circular path. To analyze this motion
we must define displacement, velocity and acceleration for this particular type of motion.
A body that moves in a circular path with a velocity whose magnitude is constant is
said to undergo uniform circular motion. Although the velocity of a body in uniform
circular motion is constant in magnitude, its direction changes continually. The body
is accelerated; the direction of this centripetal acceleration aN (normal acceleration) is
toward the center of the circle in which the body moves, and its magnitude is
v2
aN = . (2.36)
r
Here v is the speed of a particle and r is the radius of the circular path. In a circular
motion, the acceleration always changes the direction but always towards the center of
circular path.
When a particle moves in a curved path with no-constant speed, it has a tangible
acceleration (aτ ) which is along the velocity vector of the particle and normal acceleration
(aN ) which is always directed towards the centre of curvature of the curved path.
Figure 2.11: The motion of a particle along an arbitrary curved path lying in the xy plane
The magnitudes of the tangential and normal acceleration are given respectively by:
dv
aτ = (2.37a)
dt
v2
aN = (2.37b)
ρ
where ρ is the radius of curvature of the curved path at a given instant. The total
acceleration ⃗a and its magnitude are given by
q
⃗a = ⃗aN + ⃗aτ =⇒ a = a2N + a2τ (2.38)
The position of a particle can be expressed by the distance traveled s, or by the angle θ
1 1
s = s0 + v0 t + aτ t2 and θ = θ0 + ω0 t + α t2 , (2.39)
2 2
where at , ω and α are the tangential acceleration, angular velocity and angular acceleration,
respectively. The angle θ in radians is equal to the ratio between the arc s and the radius
r of the circle:
s = θr, (2.40)
and by the same comparison,
v = ωr and aτ = αr (2.41)
31
32 Lecture Notes of Physics for Engineers I Module
On the other hand, a stationary observer B will see the path of the ball as a parabola, as
illustrated in Figure 2.12.b.
In ore general situation, consider a particle located at the point P in figure Figure 2.13.
Imagine that the motion of this particle is being described by two observers, one in
′
reference frame S, fixed with respect to the earth, and another n reference S , moving to
the right relative to S with constant velocity ⃗u.
′
Figure 2.13: Fixed frame of reference, S, and the frame of reference, S , in motion with a
constant velocity, ⃗u with respect to S
We label the position of the particle with respect to the S frame with the position vector
′ ′
⃗r and label its position relative to S frame with the vector ⃗r , at some time t. If the
′
origins of the two reference frames coincide at t = 0, then the vectors ⃗r and ⃗r are related
′
to each other through expression ⃗r = ⃗r + ⃗ut or
′
⃗r = ⃗r − ⃗ut (2.44)
If we differentiate Equation 2.44 with respect to time and note that ⃗u is constant, we get
′
d⃗r ⃗r ′
= − ⃗u ⇒ ⃗v = ⃗v − ⃗u (2.45)
dt dt
Department of Physics, 2019-2020
Section 2.2: Motion in Two Dimensions 33
′
where ⃗v is the velocity of the particle observed in the S’ frame, and ⃗v is the velocity of
the same particle observed in the S frame. Equations () and () are known as Galilean
transformation equations.
Although observers in the two different references will measure different velocities for the
particle, they will measure the same acceleration when ⃗u is constant:
′
⃗v d⃗v d⃗u ′ d⃗u ⃗
= − ⇔ ⃗a = ⃗a, since =0 (2.46)
dt dt dt dt
That is, ”the acceleration of the particle measured by observer in the earth’s
frame of reference will be the same as that measured by any other observer
moving with constant velocity with respect to the first observer”.
33
Chapter 3
DYNAMICS OF A PARTICLE
Dynamics is the part of classical mechanics concerned with forces that produce changes
in motion or changes in other properties, such as the shape and the size of objects. This
leads the dynamics to the concepts of the force and the mass and the laws that govern
the motion of an object. Thus, Dynamics study the motion of objects and the causes and
effects of that motion.
34
Section 3.1: Newton’s Laws of Motion 35
Free-body diagrams are useful in identifying the forces acting on the object being considered.
Newton’s third law is also frequently needed in equilibrium problems. The two forces in
an action–reaction pair never act on the same object.
35
36 Lecture Notes of Physics for Engineers I Module
Figure 3.6: Free-body diagrams for the application of Newton’s Second Law
f k ≤ µs n (3.6)
The actual static-friction force may be anything from zero to the maximum value given
by the equation 3.6, depend on the situation. Usually µk is less than µs for a given pair
of surfaces.
When forces act on a solid object, the object usually deforms. In some cases, such as
a stretched or compressed spring, the deformation is approximately proportional to the
magnitude of the applied force. This proportionality is called Hooke’s law:
Figure 3.9: Total force acting on a particle having a non uniform circular motion
v2
arad = (3.8)
R
The period T is the time required for the object to make one complete circle; the magnitude
of the acceleration vector can also be written as
4π 2 R
arad = (3.9)
T2
Provided that the object can be treated as a particle, circular motion (like other motions)
P→ −
is governed by Newton’s second law, F = m→ −a . Because the net force vector points
in the same direction as the acceleration vector, an object in uniform circular motion is
acted upon by a net force directed toward the center of the circle, with magnitude
v2
Fnet = m (3.10)
R
In previous discussion we found that if a particle moves with varying speed in a circular
path there is, in addition of the centripetal component of acceleration, a tangential
component of magnitude || d⃗ v
dt
||. Therefore, the force acting on the particle must also
have tangential and radial components. That is, since the total acceleration is given by
⃗a = ⃗ar + ⃗aτ , the total force is given by F⃗ = F⃗r + F⃗τ , as shown in Figure 3.9.
Two particles with masses m1 and m2 a distance r apart, attract each other gravitationally
with forces of magnitude
m1 m2
Fg = G 2 (3.11)
r
These forces form an action–reaction pair in accordance with Newton’s third law. If the
objects cannot be treated as particles, but are spherically symmetric, this law is still valid;
then, r is the distance between their centers. The gravitational interaction of a spherically
37
38 Lecture Notes of Physics for Engineers I Module
Escape Speed
If an object is projected upward from Earth’s surface with a large enough speed, it can
soar off into space and never return. This speed is called Earth’s escape speed and is given
by r
2GmE
vesc = (3.16)
RE
a relation found by applying the law of conservation energy.
Kepler’s Laws
The German astronomer Johannes Kepler discovered that the orbit of Mars could be
accurately described by an ellipse with the Sun at one focus. He then generalized this
analysis to include the motions of all planets. The complete analysis is summarized in
three statements known as Kepler’s laws:
• All planets move in elliptical orbits with the Sun at one of the focal points.
• A line drawn from the Sun to any planet sweeps out equal areas in equal time
intervals.
• The square of the orbital period of any planet is proportional to the cube of the
average distance from the planet to the Sun.
The work done by the constant force is defined as the product of the component
of the force in the direction of the displacement and the magnitude of the
displacement.
Since the component of F⃗ in the direction of ⃗s is F cos θ, the work done by F⃗ is given by
W = F⃗ · ⃗s = (F cos θ) s (3.17)
According to this definition, work is done by the force F⃗ on an object under the following
conditions:
☛ the vector force F⃗ must have a non zero component in the direction of ⃗s
• From the first condition, we see that a force does no work on an object is the object does
not move (s=0).
• From the second condition, not that the work done by a force is also zero when the
force is perpendicular to the displacement, since cos θ = 0, when θ = 90o .
39
40 Lecture Notes of Physics for Engineers I Module
For example, in Figure 3.12, both the work done by the normal force and by the force of
gravity are zero since both forces are perpendicular to the displacement and have zero
component in the direction of ⃗s.
Figure 3.12: Force acting on an object which is displaced horizontally on a rough surface
⋆ The work done by the applied force is positive when the vector associated with the
component F cos θ is in same direction as the displacement. But in general, the work
is positive when done by the system and it is negative when it is done on the system.
⋆ When the vector associated with the component F cos θ is in the direction opposite
to the displacement, the work W is negative.
A common example in which the work is negative is the work done by the frictional force
when a body slides over a rough surface. If the force of sliding friction is f⃗, and the body
undergoes a linear displacement ⃗s, the work done by the force f⃗ is
Wf = −f s (3.18)
where the negative sign comes from the fact that θ = 180o and cos 180o = −1.
Finally, if an applied force F⃗ acts along the direction of the displacement, the angle θ = 0o ,
and cos 0o = 1. In this case, Equation 3.17 gives
W = Fs (3.19)
Work is a scalar quantity, and its units are force multiplied by length. Therefore, the SI
unit of force is N · m (newton × meter) which is called the joule (J). The unit of work in
the CGS system is dyne · cm, which is also called erg. Note that 1 J = 107 ergs.
∆W = Fx ∆x (3.20)
Figure 3.13: The work done by the variable force Fx as the particle moves from xi to xf
is exactly equal to the area under the curve
If we imagine that the Fx versus x curve is divided into a large number of such intervals, as
in Figure 3.13.a, the total work done for the displacement from xi to xf is approximately
xf
P
equal to the sum W ≈ Fx ∆x. If the displacements ∆x are allowed to approach
xi
zero, the value of that sum approaches a definite value equal to the true area under the
curve bounded by Fx and the x axis. This limit of the sum is called a definite integral:
xf Z xf
X
lim Fx ∆x = Fx dx
∆x→0 xi
xi
Therefore, we can express the work done by the force Fx for the displacement of the object
from xi to xf as
Z xf
W = Fx dx (3.21)
xi
W = Fx s = (max )s (3.24)
41
42 Lecture Notes of Physics for Engineers I Module
images/work_spring.png
1 1
W = mvf2 − mvi2 (3.25)
2 2
”The product of one half the mass and the square of the speed is defined as
the kinetic energy of the particle”. That is, the kinetic energy, K, of a particle of
mass m and speed v is defined as
1
K = mv 2 (3.26)
2
With this expression, it is often convenient to write Equation 3.25 as
W = ∆K = Kf − Ki (3.27)
That is, ”the work done by the resultant constant force F⃗ in displacing a
particle equals the change in kinetic energy of the particle”.
The work-energy theorem given by Equation 3.27 is also valid in the more general case
when the force varies in direction and magnitude while the particle moves along an
arbitrary curved path in three dimensions. In this situation, we express the work as
Z xf
W = F⃗ · d⃗s (3.28)
xi
where the limits i and f represent the initial and final coordinates of the particle. The
integral given by Equation 3.28 is called a line integral.
Because the infinitesimal displacement vector can be expressed as d⃗s = dx⃗i + dy⃗j + dz⃗k
and since F⃗ = Fx⃗i + Fy⃗j + Fz⃗k, Equation 3.28 reduces to
Z xf Z yf Z zf
W = Fx dx + Fy dy + Fz dz (3.29)
xi yi zi
Thus, we conclude that ”the work done on a particle, by the resultant force
acting on it, is equal to the change in the kinetic energy of the particle”.
3.3.5 Power
Power is defined as the time rate of energy transfer. If an external force is applied to an
object, and if the work done by this force is ∆W in the time interval ∆t, then the average
power during this interval is defined as the ratio of the work done to the time interval:
∆W
P = (3.30)
∆t
The instantaneous power, P , is the limiting value of the average power as ∆t approaches
zero:
∆W dW
P = lim = (3.31)
∆t→0 ∆t dt
From Equation 3.31, we can express the work done by a force F⃗ for a displacement d⃗s ,
since dW = F⃗ · d⃗s. Therefore, the instantaneous power can be written
dW d⃗s
P = = F⃗ · = F⃗ · ⃗v (3.32)
dt dt
The unit of the power in the SI system is J/s, which is also called a watt , W : 1 W =
1 J · s1 = 1 kg · m2 · s3
The unit of power in the British Engineering system is the horse power (hp), where
1hp=736 W .
43
44 Lecture Notes of Physics for Engineers I Module
With reference to the arbitrary paths shown in Figure 3.15.a, we can write the previous
condition as
WP Q (along 1) = WP Q (along 2)
A conservative force has another property, which can be derived from the above condition.
From the Figure 3.15.b, we can write
images/conservation_energy.png
WP Q (along 1) ̸= WP Q (along 2)
From this condition, we can show that if a force is nonconservative, the work done by that
force on a particle that moves through any closed path is not necessarily zero. Since the
work done in going from P to Q along path 1 is equal to the negative of the work done in
going from Q to P along path 2, it follows from the first condition of a nonconservative
force that
45
46 Lecture Notes of Physics for Engineers I Module
As an instructive example, suppose you were to displace a book between two points on a
rough horizontal surface such as a table. If the book is displaced in a straight line between
two points A and B in Figure 3.16, the work done by friction force is simply –f d, where d
is the distance between the points.
images/work_friction.png
Figure 3.16: The work done by the force of friction depends on the path taken between A
and B
However, if the book is moved along any other path between the two points, the work done
by the friction force would be greater (in absolute magnitude) than –f d. For example,
the work done by friction force along the semicircular path in Figure 3.16 is equal to
πd
−f 2 , where d is now the diameter of the circle. Finally, if the book is moved trough
any closed path (such as a circle), the work done by friction would clearly be nonzero
since the frictional force opposes the motion.
That is, the work done by a conservative force equals the negative of the change in the
potential energy associated with that force, where the change in the potential energy is
defined as ∆U = (Uf − Ui ). We can now express Equation 3.33 as
Z xf
∆U = Uf − Ui = − Fx dx (3.34)
xi
Wc = ∆K
Since the force is conservative, according to Equation 3.33 we can write thatWc = −∆U .
Hence,
Ki + U i = Kf + U f (3.36)
If we now define the total mechanical energy of the system, E , as the sum of the kinetic
and potential energies, we can express the conservation of the mechanical energy as
Ei = Ef (3.37)
where
E =K +U (3.38)
The law of conservation of mechanical energy states that ”the total mechanical
energy of a system remains constant if the only force that does work is
a conservative force”. This is equivalent to the statement that ”if the kinetic
energy of a conservative system increase (or decrease) by some amount, the
potential energy must decrease (or increase) by the same amount”.
If more than one conservative force acts on the system, then there is a potential energy
function associated with each force. In such case, we can write the law of conservation of
mechanical energy as
X X
Ki + U i = Kf + Uf (3.39)
where the number of terms in the sum equals the number of conservative forces present.
47
48 Lecture Notes of Physics for Engineers I Module
images/gravitatonalPE.png
Figure 3.17: The net work done by the force of gravity depends only on the vertical
displacements
broken down into a series of horizontal and vertical steps. There is no work done by the
force of gravity along the horizontal steps, since mg is perpendicular to these elements of
displacement. Work is done by the force of gravity only along the vertical displacements,
where the work done in the nth vertical step is −mg = ∆yn . Thus the total work done
by the force of gravity as the particle is displaced upward a distance h is the sum of the
works done along each vertical displacement. Summation of all such terms gives
X
Wg = −mg ∆yn = −mg
Since h = yf − yi , we can express Wg as
Wg = mgy (3.41)
where we have chosen to take Ug = 0 at y = 0.
Substituting the definition of Ug Equation 3.41 into the expression for the work done by
the force of gravity (Equation 3.40) gives
Wg = Ui − Uf = −∆Ug (3.42)
That is, the work done by the force of gravity is equal to the initial value of the potential
energy minus the final value of the potential energy.
The term potential energy implies that the object has the potential, or capability, of
gaining kinetic energy or doing work when released from some point under the influence of
gravity. It is often convenient to choose the surface of the earth as the reference position:
yi = 0
If the force of gravity is the only force acting on a body, then the total mechanical energy
of the body is conserved (Equation 3.37). Therefore, the law of conservation of mechanical
energy for a freely falling body can be written
1 2 1
mvi + myi = mvf2 + myf (3.43)
2 2
Wnc + Wc = ∆K.
49
50 Lecture Notes of Physics for Engineers I Module
images/potential_energy_spring.png
Now suppose there are nonconservative forces acting on the block-spring system. In this
case, we can apply the work-energy theorem in the form of Equation 3.45, which gives
1 2 1 2 1 2 1 2
E= mv + kx − mv + kx = ∆E (3.49)
2 f 2 f 2 i 2 i
That is, the total mechanical energy is not a constant of motion when non-
conservative forces act on the system. Again if Wnc is due to a force of friction,
then Wnc is negative; therefore the final energy is less than the initial energy.
p⃗ = m⃗v (3.51)
If a particle is moving in an arbitrary direction, p⃗ will have three components and
Equation 3.51 is equivalent to the component equations given by
px = mvx
py = mvy (3.52)
pz = mvz
51
52 Lecture Notes of Physics for Engineers I Module
We can relate the linear momentum to the force acting on the particle using Newton’s
second law of motion: ”The time rate of the change of the linear momentum of
a particle is equal to the resultant force on particle”. That is
⃗ ∆→−
p d⃗p
F = lim = (3.53)
∆t→0 ∆t dt
From Equation Equation 3.53 we see that if the resultant force is zero, the momentum
of the particle must be constant. In other words, the linear momentum of a particle is
conserved when F⃗ = 0. That is, when the force is zero, the acceleration of the particle is
zero and the velocity remains constant.
Equation 3.53 can be written
d⃗p = F⃗ dt (3.54)
If the momentum vector of the particle changes from p⃗i at time ti to p⃗f at time tf , then
integrating Equation 3.54 gives
Z tf
∆⃗p = p⃗f − p⃗i = F⃗ dt (3.55)
ti
The quantity on the right side of Equation 3.55 is called the impulse vector of the
force F⃗ for the time interval ∆t = tf − ti . Impulse is the vector defined by
Z tf
J⃗ = F⃗ dt = ∆⃗p (3.56)
ti
That is, ”the impulse vector of the force F equals the change in the momentum
vector of the particle”. This statement, known as the impulse-momentum theorem, is
equivalent to Newton’s second law of motion. From this definition, we see that impulse is
a vector quantity having a magnitude equal to the area under the force–time curve, as
described in Figure 3.19.a.
Since the force can generally vary in time as in Figure 3.19.a, it is convenient to define a
time averaged force F⃗ , given by
Z tf
⃗ = 1
F F⃗ dt = ∆⃗p (3.57)
∆t ti
⃗ ∆t
J⃗ = F (3.58)
This average force, described in Figure 3.19.b, can be thought of as the constant force that
would give the same impulse to the particle in the time interval ∆t as the time-varying
force gives over this same interval.
In principle, if F⃗ is known as a function of time, the impulse can be calculated from
Equation 3.56. The calculation becomes especially simple if the force acting on the particle
⃗ = F⃗ and Equation 3.58 becomes
is constant. In this case, F
J⃗ = ∆⃗p = F⃗ ∆t (3.59)
images/impulse.png
Figure 3.19: The impulse is the area under the force versus curve
3.5.2 Conservation of Linear Momentum for a Two-Particles
System
Consider two particles that can interact with each other but are isolated from their
surroundings Figure 3.20. That is, the particles exert forces each other, but no external
forces are present. Suppose that at some time t, the momentum of particle 1 is p1 and the
momentum of particle 2 is p2 . We can apply Newton’s second law to each particle and write
F⃗12 = d⃗
p1
dt
and F⃗21 = d⃗p2
dt
where F⃗12 is the force on particle 1 due to the particle 2, and F⃗21 is the force on particle 2
due to particle 1.
However, Newton’s third law of motion tells us that F⃗12 and F⃗21 are equal in magnitude
and opposite in direction. That us, they form an action-reaction pair and F⃗12 = −F⃗21 .
we can express this condition as
d⃗p1 d⃗p2 d
F⃗12 + F⃗21 = 0 or + = (⃗p1 + p⃗2 ) = 0
dt dt dt
Since the time derivative of the total momentum, p⃗ = p⃗1 + p⃗2 , is zero, we conclude that
the total momentum, p⃗ must remain constant, that is
This vector equation is equivalent to three component equations. In other words, Equa-
tion 3.60 in component form says that the total momenta in the x, y, and z directions
are all independently conserved, or
53
54 Lecture Notes of Physics for Engineers I Module
images/Lmonentum_conserv.png
Figure 3.20: The total momentum of the system ,⃗p, is equal to the vector sum p⃗1 + p⃗2 and
F⃗12 = −F⃗21
We can state this law, known as the conservation of linear momentum, as follow: ”if
two particles of massesm1 and m2 form an isolated system, then the total
momentum of the system is conserved, regardless of the nature of the force
between them”. More simply, whenever two particles collide their total momentum
remains constant, provided they are isolated.
Suppose that ⃗v1i and ⃗v2i are the initial velocities of particles 1 and 2, and ⃗v1f and ⃗v2f
are their velocities at some later time. Applying Equation 3.60, we can express the
conservation of linear momentum of this isolated system in the form
Figure 3.22: Collisions between two cars as the result of direct contact
When two particles of masses m 1 and m 2 collide as in Figure 3.22, the impulse forces
may vary in time in a complicated way such as described in Figure 3.23.
If F⃗12 is the force on the mass m1 due to m2 then the change in momentum of m1 due to
the collision is given by
Z tf
∆⃗p1 = F⃗12 dt
ti
If F⃗21 is the force on the mass m2 due to m1 then the change in momentum of m2 due to
the collision is given by
Z tf
∆⃗p2 = F⃗21 dt
ti
However, Newton’s third law states that the force on mass m1 due to m2 is equal and
opposite the force on mass m2 due to m1 , or F⃗12 = F⃗21 (see Figure 3.23). Hence, we
conclude that
55
56 Lecture Notes of Physics for Engineers I Module
Whenever a collision occurs between two bodies, we have seen that the total momentum
is always conserved . However, the total kinetic energy is generally not conserved
when collision occurs because some of kinetic energy is converted into thermal
energy and internal inelastic potential energy when the bodies are deformed during
the collision.
⋆ We define an inelastic collision as a collision in which momentum is conserved
but kinetic energy is not.
⋆ When two objects collide and stick together after the collision, the collision is called
perfectly inelastic.
This is an extreme case of an inelastic collision. For example, if two pieces of putty collide,
they stick together and move with some common velocity after the collision.
⋆ An elastic collision is defined as a collision in which both momentum and kinetic
energy are conserved . Billiard ball collisions and the collision of air molecules with the
walls of a container at ordinary temperatures are highly elastic. Elastic and perfectly
inelastic are the limiting cases, and most collisions are cases in between.
Mathematically, a coefficient known as Coefficient of restitution is used to classify collisions,
with
vB,f − vA,f
e= (3.63)
vA,i − vB,i
where vA,i and vB,i are the velocities of mass A and B before impact, and vA,f and vB,f
are the velocities of mass A and B after impact (see Figure 3.24).
For a perfectly elastic collisions, e = 1. For inelastic collisions e < 1. For the bodies stick
together after collision (perfectly inelastic collisions), e = 0.
If two particles stick together and move with some common velocity ⃗vf after collision,
then only the linear momentum of the system is conserved:
m1⃗v1i + m2⃗v2i
⇒ ⃗vf = (3.65)
m1 + m2
In this case, both momentum and kinetic energy are conserved, therefore we can write
these conditions:
2 2 2 2
m1 (v1i − v1f ) = m2 (v2f − v2i )
Next, let us factor both sides of the equation to get
☛ If m1 = m2 then we see that v1f = v2i and v2f = v1i . That is, the particles
exchange velocities if they have equal masses; this is one observes in billiard ball
collision.
☛ If m1 = 0 is initially at rest, v2i = 0 and equations Equation 3.70 becomes
(m1 − m2 )v1i
v1f = (3.71a)
m1 + m2
2m1 v1i
v2f = (3.71b)
m1 + m2
57
58 Lecture Notes of Physics for Engineers I Module
☛ If m1 is very large compared with m2 and v2i = 0, we see from Equation 3.71
that v1f ≈ v1i and v2f ≈ 2v1i . That is, when a very heavy particle collides head-on
with a very light one initially at rest, the heavy particle continues its motion
unaltered after the collision, while the light particle rebounds with a
velocity equal to about twice the initial velocity of the heavy particle.
After the collision, m1 moves at an angle θ with respect to the horizontal and m2 moves at
an angle ϕ with respect to the horizontal. Applying the law of conservation of momentum
in component form, pxi = pxf and pyi = pyf and noting that pyi = 0, we get
Now let us assume that the collision is elastic, in which case we can also write a third
equation for the conservation of kinetic energy, in the form
1 2 1 2 1 2
m1 v1i = m1 v1f + m2 v2f (3.73)
2 2 2
If we know the initial velocity, ⃗v1i , and the masses, we are left with four unknowns. Since
we only have three equations, one of the remaining quantities (v1f , v2f , θ, or ϕ) must
be given to determine the motion after the collision from conservation principles alone.
DYNAMICS OF A SYSTEM OF
PARTICLES
In this section we describe the overall motion of a mechanical system in terms of a very
special point called the center of mass of a system. The mechanical system can be either a
system of particles or an extended object. We shall see that mechanical system moves as if
all its mass were concentrated at the center of mass. Furthermore, if the resultant external
force on the system is F⃗ and the total mass of the system is M , the center of mass moves
⃗
F
with an acceleration given by ⃗a = M That is, the system moves as the resultant external
force were applied to a single particle of mass M located at the center of mass.
One can describe the position of the center of mass of a system as being the
average position of the system’s mass. For example, the center of mass of the pair
of particles described in Figure 4.1 is located on the x axis and lies somewhere between
the particles.
m1 x1 + m2 x2
xc = (4.1)
m1 + m2
For example, if x1 = 0, x2 = d, and m2 = 2m1 , we find that xc = 23 d That is, the center
of mass lies closer to the more massive particle. If the two masses are equal, the center of
mass lies midway between the particles.
We can extend the center of mass concept to a system of many particles in three dimensions.
The x, y and z coordinates of the center of mass of n particles is defined to be
59
60 Lecture Notes of Physics for Engineers I Module
n
P
mi xi
i=1
xc = Pn (4.2a)
mi
i=1
n
P
mi yi
i=1
yc = Pn (4.2b)
mi
i=1
n
P
mi zi
i=1
zc = Pn (4.2c)
mi
i=1
n
P
For convenience, we shall express the total mass as M = mi
i=1
The center of mass can also be indicated by its position vector,⃗rc . The rectangular
coordinates of this vector are xc , yc , and zc defined in equations Equation 4.2. Therefore,
n
P n
P n
P
mi xi + mi yi + mi zi
i=1 i=1 i=1
⇒ ⃗rc = (4.3)
M
or
n
P
mi⃗ri
i=1
⃗rc = (4.4)
M
where ⃗ri is the position of the ith particle, defined by ⃗ri = xi⃗i + yi⃗j + zi⃗k
If the body is considered to have a continuous mass distribution, we can express the
vector position of the center of mass of a rigid body in the form
Z
1
⃗rc = ⃗rdm (4.5)
M
where this is equivalent to the three scalar expressions given by the following equation
Z
1
xc = xdm (4.6a)
M
Z
1
yc = ydm (4.6b)
M
Z
1
zc = zdm (4.6c)
M
where ⃗vi is the velocity of the ith particle. Rearranging Equation 4.7 yields
n
X n
X
M⃗vc = mi⃗vi = p⃗i = P⃗ (4.8)
i=1 i=1
We conclude that the total momentum vector of the system equals the total
mass multiplied by the velocity vector of the center of mass. In other words,
the total momentum of the system is equal to that of a single particle of mass M moving
with velocity vector ⃗vc .
If we now differentiate Equation 4.7 with respect to time, we get the acceleration vector
of the center of mass
n n
d⃗vc 1 X d⃗vi 1 X
⃗ac = = mi = mi⃗ai (4.9)
dt M i=1 dt M i=1
Rearranging this expression (Equation 4.9) and using Newton’s second law, we get
n
X n
X
M⃗ac = mi⃗ai = F⃗i (4.10)
i=1 i=1
61
62 Lecture Notes of Physics for Engineers I Module
dP⃗
= M⃗ac = ⃗0, so that P⃗ = M⃗vc = Constant (4.12)
dt
For an isolated system of particles, both the total momentum and velocity of the center
of mass are constant in time.
Figure 4.2: Rotation of a rigid body about a fixed axis through O perpendicular to the
plane of the figure (z axis)
Figure 4.2 illustrates a planar rigid body of arbitrary shape confined to the xy plane and
rotating about a fixed axis through O perpendicular to the plane of the figure. A particle
on the body at a point P is at a fixed distance r from the origin and rotates in a circle of
radius r about O. It is convenient to represent the position of the point P with its polar
coordinates (r, θ). As the particle moves along the circle from the positive x axis θ = 0 to
the point P, it moves through an arc length s, which is related to the angular positionθ
through the relation
s
s = rθ ⇒ θ= (4.13)
r
The quantity θ is then a pure number since it is the ratio of an arc length and the radius
of a circle.
As the particle travels from P to Q in the Figure 4.3 in a time ∆t, the radius sweeps out
an angle ∆θ = θ2 − θ1 , which equals the angular displacement.
Figure 4.3: A particle on a rotating rigid body moving from P to Q along the arc of a
circle
We define the average angular velocity ω as the ratio of this angular displacement to the
time interval ∆t:
θ2 − θ1 ∆θ
ω= = (4.14)
t2 − t1 ∆t
In analogy to the linear velocity, the instantaneous angular velocity, ω, is defined as the
limit of the ratio in Equation 4.14 as ∆t approaches zero:
∆ω dω
ω = lim = (4.15)
∆t→0 ∆t dt
If the instantaneous angular velocity of a body changes from ω1 to ω2 in the time interval
∆t, the body has an angular acceleration. The average angular acceleration, α, of a
rotating body is defined as the ratio of the change in the angular velocity to the time
interval ∆t:
ω2 − ω1 ∆ω
α= = (4.16)
t2 − t1 ∆t
In analogy to linear acceleration, the instantaneous angular acceleration, α, is defined as
the limit of the ratio as ∆ω
∆t
approaches zero:
∆ω dω
α = lim = (4.17)
∆t→0 ∆t dt
For rotation about a fixed axis, we see that every particle on rigid body has the same
angular velocity and the same angular acceleration. That is, the quantities ω and α
characterize the rotational motion of the entire rigid body.
We shall take ω to be positive when s increasing (counterclockwise motion) and negative
when θ is decreasing (clockwise motion). The direction of ω
⃗ is along the axis of rotation,
which the z axis in Figure 4.2. By convention, we take the direction of ω ⃗ to be out of
the plane of the diagram when the rotation is counterclockwise, and into the plane of the
diagram when the rotation is clockwise.
Finally, the sense of α follows from its definition as dω
dt
. It is the same as ω if the angular
dt speed ω is increasing in time and anti parallel to ω ⃗ if the angular speed is decreasing
in time.
ω = ω0 + αt (f or α = Constant) (4.18)
Likewise, substituting Equation 4.18 into Equation 4.17 and integrating once more (with
θ = θo at to = 0), we get
1
θ = θo + ω0 t + αt2 (4.19)
2
63
64 Lecture Notes of Physics for Engineers I Module
Notice that these kinematic expressions for rotational motion under constant angular accel-
eration are of the same form as those for linear motion under constant linear acceleration
with substitutions:
x → θ
v→ω (4.21)
a→α
Figure 4.4: The point P has a linear velocity ⃗v , always tangent to the circular path of the
radius r
We first relate the angular velocity of the rotating body to the tangential velocity, ⃗v , of
a point P on the body. Since P moves in a circle, the linear velocity vector is always
tangent to the circular path, and hence the phrase tangential velocity. Recalling that
s = rθnd noting that r = Constant, we get
ds dθ
v= =r ⇒ v = rω (4.22)
dt dt
Figure 4.5: Total acceleration of the point P of a rigid body rotating about a fixed axis
through O
v2
ar = = rω 2 (4.23)
r
The total linear acceleration of the particle is ⃗a = ⃗at + ⃗ar . Therefore, the magnitude of
the total linear acceleration of the point P on the rotating rigid body is given by
q √ √
a = a2t + a2r = r2 α2 + r2 ω 2 = r α2 + ω 2 (4.24)
Figure 4.6: Kinetic energy of the particle of mass mi of a rigid body rotating about z axis
with angular velocity ω
4.3.4 Rotational Kinetic Energy
Let us consider a rigid body as a collection of small particles and let us assume that the
body rotates about the fixed z axis with an angular velocity ω (Figure 4.6).
If the mass of the particle is mi and its speed is vi , the kinetic energy of this particle is
1
Ki = mi vi2
2
We recall that although every particle in the rigid body has the same angular velocity
ω , the individual linear velocities depend on the distance ri from the axis of rotation
according to the expression vi = ri ω . The total kinetic energy of the rotating rigid body
is the sum of the kinetic energies of the individual particles:
!
X 1X 1X 1 X
K= Ki = mi vi2 = mi ri2 ω 2 ⇒ K= mi ri2 ω2 (4.25)
2 i 2 i 2 i
Using this notation, we can express the kinetic energy of the rotating rigid body (Equa-
tion 4.25) as
1
K = Iω 2 (4.27)
2
That is, the rotational analogy of mass is a quantity called moment of inertia. The
greater the moment of inertia of a body, the greater its resistance to change its angular
velocity.
When the mass of the body is uniformly (continuously) distributed, the sum is reduced to
the integral as: Z
I = r2 dm
where the integral is extended over all the volume of the solid. In this case, the quantity
r is a function of the coordinates x, y and z. Figure 4.7 gives the moments of inertia of
some homogeneous bodies of mass m.
65
66 Lecture Notes of Physics for Engineers I Module
I = Ic + ma2 (4.29)
Radius of gyration
The radius of gyration of a body is the quantity K defined as followed:
r
I
I = mK 2 or K = (4.30)
m
with m as the mass of a body and I its moment of inertia about an axis of rotation. The
radius of gyration represents a radial distance from any given axis at which the mass of
the body could be concentrated without altering the moment of inertia of the body about
that axis.
4.3.6 Torque
When a force is exerted on a rigid body pivoted about some axis, the body will tend
to rotate about that axis. The tendency of a force to rotate a body about some axis is
measured by a quantity called the torque (⃗τ ). Consider the wrench pivoted about the
axis through O in Figure 4.8. The applied force F⃗ generally can act at an angle ϕ to the
horizontal.
Figure 4.8: The force F⃗ gas greater tendency to rotate about O as F⃗ increases and d
increases
We define the magnitude of the torque, τ , resulting from the force F⃗ by the expression.
τ = rF sin ϕ = F d (4.31)
It is very important to recognize that torque is defined only when a reference axis is
specified. The quantity d rF sin ϕ, called the moment arm (or lever arm) of the force
F⃗ , represents the perpendicular distance from the rotation axis to the line of action of
F⃗ . Note that the only component of F⃗ that tends to cause a rotation is F⃗ sin ϕ, the
component perpendicular to ⃗r. The horizontal component, F⃗ cos ϕ, passes through O and
has no tendency to produce a rotation.
Thus, the torque is a vector quantity which is given by
⃗τ = ⃗r × F⃗ (4.32)
Figure 4.9: The force F⃗2 tends to rotate the body counterclockwise about O, and tends to
rotate the body clockwise
For example, in Figure 4.9, the torque τ1 resulting from F⃗1 , which has a moment arm d1 ,
is positive and equal to +F1 d1 , the torque τ2 resulting from F⃗2 , which has a moment arm
d2 , is negative and equal to −F2 d2
Hence, the net torque acting on the rigid body about O is
τnet = τ1 + τ2 = F1 d1 − F2 d2 (4.33)
Figure 4.10: A particle rotating in a circle under the influence of a tangential force F⃗t
τ = Ft r = (mat )r
Since the tangential acceleration is related to the angular acceleration through the relation
at = rα the torque can be expressed as
τ = Iα (4.34)
That is, the torque acting on the particle is proportional to its angular ac-
celeration, and the proportionality constant is the moment of inertia. It
is important to note that τ = Iα is the rotational analogue of Newton’s second law of
motion, F = ma.
67
68 Lecture Notes of Physics for Engineers I Module
Figure 4.11: A rigid body rotating about an axis through O under the action of an external
force F⃗ applied at P
Since the magnitude of the torque due to F⃗ about the origin was defined as rF sin ϕ, we
can write the work done for the infinitesimal rotation dθ as
dW = τ dθ (4.36)
The rate at which work is being done by F⃗ for rotation about the fixed axis is obtained
by formally dividing the left and right sides of Equation 4.36 by dt :
dW θ
= τd (4.37)
dt dt
But the quantity dW
dt
is, by definition, the instantaneous power, P , delivered by the force.
Furthermore, since dtθ = τ ω, Equation 4.37 reduces to
dW θ
P = = τd (4.38)
dt dt
The expression in analogous to P = F v in the case of linear motion, and the expression
dW = τ dθ is analogous to dW = Fx dx
τ dθ = dW = Iωdω
Integrating this expression, we get for the total work done
Z θ Z θ
1 1
W = τ dθ = Iωdω = Iω 2 − Iωo2 = ∆Krot (4.39)
θo θo 2 2
That is, ”the net work done by external forces in rotating a symmetric rigid
body about a fixed axis equals the change in the body’s rotational kinetic
energy”.
analogy to the conservation of linear momentum, we shall find that the angular
momentum of any isolated system is always conserved .
This conservation law is a special case of the result that the time rate of the change of
the total angular momentum of any system of particles equals the resultant
external torque acting on the system.
Figure 4.12: Paths taken by the c.m. and the point on the rim are different
The center of mass moves in a straight line, while a point on the rim moves in more
complex path which corresponds to the path of a cycloid.
Now consider a uniform cylinder of radius R rolling on a rough, horizontal surface as in
Figure 4.13. Therefore, the velocity and acceleration of the center of mass for pure rolling
motion are given by
ds dθ
vc = = R = Rω (4.40)
dt dt
vc dω
ac = =R = Rα (4.41)
dt dt
Figure 4.13: For pure rolling motion, as the cylinder rotates through an angle θ, the center
of mass of the cylinder moves a distance s = Rθ
The linear velocities of various points on the rolling cylinder are illustrated in Figure 4.14.
Figure 4.14: All points on the rolling body move in a direction perpendicular to an axis
through the contact point P
At any instant, the point P is at rest relative to the surface since sliding does not occur.
A general point on the cylinder, such as Q, has both horizontal and vertical components
of the velocity. However, the points P, P’, and C at the center of mass are unique and of
special interest. Relative to the surface on which the cylinder is moving, the center of
mass moves with velocity vc = Rω whereas the contact point P has zero velocity. The
point P’ has a velocity equal to 2vc = 2Rω, since all points on the cylinder have the same
angular velocity.
We can express the total kinetic energy of the rolling cylinder as
1
K = Ip ω 2 (4.42)
2
69
70 Lecture Notes of Physics for Engineers I Module
⃗ = ⃗r × p⃗
L (4.44)
⃗ is perpendicular to the plane formed by ⃗r and p⃗, and its sense is
The direction of L
⃗ is given by
governed by the right-hand rule. Since p⃗ = m⃗v , the magnitude of L
d⃗p
⃗τ = ⃗r × F⃗ = ⃗r × (4.46)
dt
Now let us differentiate Equation 4.44 with respect to time
⃗
dL d d⃗p d⃗r
= (⃗r × p⃗) = ⃗r × + × p⃗
dt dt dt dt
d⃗
r
The last term on the right hand in the above equation is zero, since ⃗v = dt
is parallel to
p⃗.
Therefore,
dL⃗ d⃗p
= ⃗r × (4.47)
dt dt
comparing equations (4.46) and (4.47), we see that
⃗
dL
τ= (4.48)
dt
which is the rotational analogy of Newton’s second law, F⃗ = d⃗
p
dt
. This result says that
“the vector torque acting on a particle is equal to the time rate of the change
of the particle’s angular momentum vector”.
Since the individual angular momentum vectors of the particles may change in time, the
total angular momentum vector may also vary in time. That is, we find that the time
rate of the change of the total angular momentum vector equals the vector sum of all
torque vectors, including those associated with internal forces between particles and those
associated with external forces. However, the net vector torque associated with the internal
forces is zero. Finally, we conclude that the total angular momentum vector can vary with
time only if there is a net external vector torque on the system, so that we have
n ⃗i n ⃗
X X dL d X⃗ dL
⃗τext = = Li = (4.50)
i=1
dt dt i=1 dt
That is, “the time rate of the change of the total angular momentum vector
of a system about some origin in an inertial frame equals the net external
vector torque acting on the system about that origin”.
⃗ is
Figure 4.16: When a rigid body rotates about an axis, the angular momentum vector L
in the same direction as the angular velocity vector ω
⃗
71
72 Lecture Notes of Physics for Engineers I Module
!
X X
Lz = mi ri2 ω = mi ri2 ω or Lz = Iω (4.51)
i i
where Lz is the component of the angular momentum and I is the moment of inertia of
the rigid body about the z axis.
Now let us differentiate Equation 4.51 with respect to time noting that I is constant for a
rigid body
dLz dω
=I = Iα (4.52)
dt dt
where α is the angular acceleration relative to the axis of rotation z . Because the
product Iα is equal to the net torque (see Equation 4.50), we can express Equation 4.52
as follows
X dLz
= Iα
τext = (4.53)
dt
That is, “the net external torque acting on a rigid body rotating about a fixed
axis equals the moment of inertia about the axis of rotation multiplied by its
angular acceleration relative to that axis”.
⃗i = L
L ⃗f = C
⃗ st (4.56)
If the system is a body rotating about a fixed axis, such as the z axis, then we can write
Lz = Iω
where Lz is the component of L⃗ along the axis of rotation; and I is the moment of inertia
about this axis.
In this case, we can express the conservation of angular momentum as
Ii ωi = If ωf = C st (4.57)
This expression is valid for rotations either about a fixed axis or about an axis through
the center of mass of the system as long as the axis remains parallel to itself.
Although we do not prove it here, there is an important theorem concerning the angular
momentum vector relative to the center of mass. This theorem states that “the resultant
vector torque acting on a body about the center of mass equals the time rate
of the change of angular momentum vector regardless of the motion of
the center of mass”. This theorem applies even if the center of mass is accelerating,
⃗ are evaluated relative to the center of mass.
provided that ⃗τ and L
73
Chapter 5
STATIC EQUILIBRIUM AND
ELASTICITY
5.1 Introduction
Part of this chapter is concerned with the conditions under which a rigid body is in
equilibrium. The term equilibrium implies that the body is either at rest or
that its center of mass moves with constant velocity . We shall deal with bodies
at rest or bodies in static equilibrium.
In dynamics we stated that one necessary condition for equilibrium is that the net vector
force on an object is zero. If the object is treated as a single particle, this is the only
condition that must be satisfied in order that the particle is in equilibrium. That is, “if
the net vector force on the particle is zero, it will remain at rest (if originally
at rest) or move with constant velocity vector in a straight line (if originally
in motion)”.
The situation with real objects is somewhat more complex because the objects cannot
be treated as particles. An object has a definite size, shape, and mass distribution. In
order for an object to be in static equilibrium, the net vector force on it must
be zero and the object must have no tendency to rotate. This second condition
of equilibrium requires that the net vector torque about any origin be zero. In order to
establish that whether or not an object is in equilibrium, we must know the size and
shape of the object, and the points of application of the various forces.
In the first part of this chapter, we shall be concerned with objects that are assumed
to be rigid. A rigid object is defined as one that does not deform under the
application of external forces. That is, all parts of a rigid object remain at a fixed
separation with respect to each other when subjected to external forces.
The last section of this chapter deals with the realistic situation of objects that deform
under load conditions. Such deformations are usually elastic in nature and will not
affect the conditions of equilibrium. By elastic we mean that when the deforming
forces are removed, the object returns to its original shape. Several elastic
constants will be defined, each corresponding to a different type of deformation.
Figure 5.1: The moment arm of F⃗ relative to O is the perpendicular distance d from O to
the line of action of F⃗
74
Section 5.2: Conditions of equilibrium of a rigid object 75
⃗τ = ⃗r × F⃗
Recall that the sense of τ is determined by the sense of rotation that F⃗ tends to give to
the object. The right-hand rule can be used to determine the direction of ⃗τ . By definition,
the magnitude of ⃗τ is given by F d.
Now suppose that two vector forces, F⃗1 and F⃗2 , act on a rigid object. The two vector
forces will have the same effect on the object only if they have the same magnitude, the
same direction, and the same line of action.
In other words, “two vector forces F⃗1 and F⃗2 are equivalent if and only if
|F⃗1 | = |F⃗2 | and if they have the same torque about any given point”. An
example of two equal and opposite forces that are not equivalent is shown in Figure 5.2.
Figure 5.2: The two vector forces acting on the object are equal in magnitude and opposite
in direction; yet the object is not in equilibrium
The vector force directed toward the right tends to rotate the object clockwise about
the axis through O, whereas the vector force directed toward the left tends to rotate it
counterclockwise about that axis.
Now suppose an object is pivoted about an axis through its center of mass as in Figure 5.3.
Two equal and opposite vector forces act in directions shown, such that their lines of
action do not pass through the center of mass. A pair of vector forces acting in this
Figure 5.3: The two vector forces acting on the object are equal in magnitude and opposite
in direction; yet the object is not in equilibrium
manner forms what is called a couple. Since each force produces the same torque,
Fd , the net vector torque has magnitude given by 2 Fd . Clearly, the object will rotate
in clockwise direction and will undergo an angular acceleration about the axis. This is
a non-equilibrium situation as far as the rotational motion is concerned. That is, the
unbalanced, or net torque on the object gives rise to an angular acceleration
α according to the relationship τnet = −2F d = Iα
In general, an object will be in rotational equilibrium only if its angular
acceleration α = 0. Since τnet = Iα or rotation about a fixed axis, a necessary condition
of equilibrium for an object is that the net torque about any origin must be zero. We now
have two necessary conditions for equilibrium of an object, which can be stated as follows:
(a) the resultant external vector force must equal a zero vector :
X
F⃗ext = ⃗0 (5.1)
(b) the resultant external vector torque must be a zero vector about any
origin: X
⃗τext = ⃗0 (5.2)
75
76 Lecture Notes of Physics for Engineers I Module
⋆ The first condition is a statement of translational equilibrium, that is, “the linear
acceleration must be zero viewed from an inertial reference frame”.
⋆ The second is a statement of rotational equilibrium, that is, “the angular acceleration
about any axis must be zero”.
In the special case of static equilibrium, which is the main subject of this chapter, the
object is at the rest so that it has no linear or angular velocity (that is, ⃗vc = ⃗0 and ω
⃗ = ⃗0)
The two vector expressions given by equations (5.1) and (5.2) are equivalent, in general
to six scalar equations. Three of these come from the first condition of equilibrium, and
three follow from the second condition (corresponding to x, y, and z components). We
will restrict our discussion to situations in which all the vector forces lie in a common
plane, which we assume to be the xy plane. Vector forces whose representations are in
the same plane are said to be coplanar. In this case, we shall have to deal with only three
scalar equations. Hence, these conditions of equilibrium provide the equations
P
P Fx = 0
Fy = 0 (5.3)
P
τz = 0
There are two cases of equilibrium that are often encountered.
Case I: If an object is subjected to two vector forces, the object is in equilib-
rium if and only if the two vector forces are equal in magnitude, opposite in
direction, and have the same line of action.
Figure 5.4a shows a situation in which the object is not in equilibrium because the two
vector forces are not along the same line of action. In Figure 5.4b, the object is in
equilibrium because the vector forces have the same line of action.
Figure 5.5: If three vector forces act on an object that is in equilibrium, their lines of
action must intersect at a point S (or they must be parallel)
The conditions of equilibrium require that F⃗1 + F⃗2 + F⃗3 = ⃗0 and that net vector torque
about any axis be a zero vector.
We can now finish this section assuming that “if any object is in translational
equilibrium and the net vector torque is zero about one point; it must be zero
about any other point”.
the rod such that the torque exerted by a single particle of mass M = m1 + m2 + m3 at
X equals the sum of the torques exerted by the particles at their locations x1 , x2 , and x3 .
Thus,
m1 gx1 + m2 gx2 + m3 gx3 = M gX = (m1 + m2 + m3 )gX
P
m 1 x1 + m 2 x2 + m 3 x3 mi xi
X= = Pi (5.4)
m1 + m2 + m3 i mi
This formula can be extended to any number of particles. If the complex object involves
two or three dimensions rather than just one, the same procedure is applied along two
or three coordinate axes to find X, Y and Z, which are the coordinates of the center of
gravity.
77
78 Lecture Notes of Physics for Engineers I Module
F⊥
T ensile Stress ≡ σ = (5.5)
A
When the forces on the ends of a bar are pushes rather than pulls (Figure 5.9a), the bar
is in compression, and the stress is a compressive stress. At the cross section shown,
each side pushes, rather than pulls, on the other.
The fractional change in length (the stretch) of an object under a tensile stress is called
the tensile strain. Figure 5.10a shows a bar with unstretched length lo that stretches to a
length l = lo + ∆l when equal and opposite forces with magnitude F⊥ are applied to its
ends. The elongation ∆l, measured along the same line as lo , doesn’t occur only at the
ends; every part of the bar stretches in the same proportion. The tensile strain is defined
as the ratio of the elongation ∆l to the original length lo :
l − lo ∆l
T ensile Stain ≡ ϵ = = (5.6)
lo lo
Tensile strain is the amount of stretch per unit length. It is a ratio of two lengths, always
measured in the same units, so strain is a pure (dimensionless) number with no units. The
compressive strain of a bar in compression is defined in the same way as tensile strain,
but ∆l has the opposite direction (Figure 5.10b). In this case, it is often convenient to
treat ∆l as a negative quantity.
F⊥ /A lo F⊥
Y = = (5.7)
∆l/lo A ∆l
Young’s modulus is a property of a specific material, rather than of any particular object
made of that material. A material with a large value of Y is relatively unstretchable; a
large stress is required for a given strain. For example, steel has a much larger value of Y
than rubber.
79
80 Lecture Notes of Physics for Engineers I Module
Volume strain is change in volume per unit volume. Like tensile or compressive strain, it
is a pure number, without units.
When Hooke’s low is obeyed, the volume strain is proportional to the volume stress
(change in pressure). The corresponding constant ratio of stress to strain is called the
bulk modulus, denoted by B When the pressure on an object changes by a small amount
∆p, from po to po + ∆p, and the resulting volume strain is ∆V /Vo , Hooke’s law takes the
form
∆p
B=− (5.10)
∆V /Vo
We include a minus sign in this equation because an increase in pressure always causes a
decrease in volume. In other words, when ∆p is positive, ∆V is negative. The negative
sign in Equation 5.10 makes B itself a positive quantity.
F∥
Shear stress = − (5.11)
A
Shear stress, like the other two types of stress, is a force per unit area. For systems in
equilibrium, shear stress can exist only in solid materials; a fluid would simply flow in
response to the tangential force.
Figure 5.12b shows an object deformed by shear. We define shear strain as the
ratio of the displacement x to the transverse dimension h; that is,
x
Shear strain = = tan ϕ (5.12)
h
with x and h defined as in Figure 5.12b. In real-life situations, x is nearly always much
smaller than h, tan ϕ is very nearly equal to ϕ, and the strain is simply the angle ϕ,
measured in radians. Like all strains, shear strain is a dimensionless number because it is
a ratio of two lengths.
If the forces are small enough so that Hooke’s law is obeyed, the shear strain is proportional
to the shear stress. The corresponding proportionality constant (ratio of shear stress to
shear strain), is called the shear modulus, denoted by S:
Shear stress F∥ /A h F∥ F∥ /A
S= = = = (5.13)
Shear strain x/h A x ϕ
For a given material, S is usually a third to a half as large as Y .
Elastic region: the material is elastic and the deformation disappears completely if the
stress is no longer acting;
81
82 Lecture Notes of Physics for Engineers I Module
Plastic-elastic region: the deformation does not disappear completely after the decay
of the stress;
Plastic region: the deformation is also maintained to a large extent without stress (i.e.
region of permanent deformation).
Hooke’s straight line: tangent to the stress-strain curve at the origin. Its slope is the
elastic modulus E of the body for small strains.
F/A Y Al
Y = , hence F =
∆l/L L
where Y is the Young’s modulus for the material of the wire. The work done for additional
small increase dl in the length of the wire is
YA
dW = F dl = ldl
L
The work done during the whole increase l in length of the wire is
1 Y Al2
Z Z
YA 1
W = dW = ldl = = Fl (5.15)
L 2 L 2
The work done during stretching of the wire=1/2 stretch f orce × stretch. Since volume
of the wire is V = A · L, the work done per unit volume of the wire or the strain energy
per unit volume is
W 1F l 1
w= = = stress × strain (5.16)
V 2AL 2
6.1 Introduction
In this chapter, first we shall present a discussion of the various states of matter .
Next, we shall consider a fluid at rest and derive an expression for the pressure as a
function of its density and depth. We shall then treat fluids in motion, or fluid dynamics.
An underlying principle known as the Bernoulli principle will enable us to determine
relations between the pressure, density, and velocity at every point in a fluid. As we shall
see, the Bernoulli principle is a result of conservation of energy applied to
an ideal fluid .
Figure 6.1: A model of a solid: the eight atoms are imagined as being attached to each
other by springs, which represent the elastic nature of the inter-atomic forces.
83
84 Lecture Notes of Physics for Engineers I Module
solid tends to return to its original shape and size. For this reason, a solid is
said to have elasticity.
In any given substance, the liquid state exists at a higher temperature than the solid state.
Thermal agitation is greater in the liquid state than in the solid state. As a result, the
molecular forces in a liquid are not strong enough to keep the molecules in fixed positions,
and the molecules wander through the liquid in random fashion (Figure 6.2). Solids and
liquids have the following property in common. When one tries to compress a liquid or a
solid, strong repulsive atomic forces act internally to resist the deformation.
In gaseous state, the molecules are in constant random motion and exert only weak forces
on each other. The average separation distances between the molecules of the gas are
quite large compared with the dimensions of the molecules. Occasionally, the molecules
collide with each other; however, most of time they move as nearly free, non interacting
particles.
• Mass Density, ρ = m
V
• Weight Density, ω = ρg
(b) Viscosity
The viscosity of a fluid is that property which determines the amount of its resistance to
a shearing force. Viscosity is due primarily to interaction between molecules.
We have two coefficients of viscosity:
F dv
τ= =µ (6.1)
A dy
where, τ is the shear stress, dv the variation of the velocity, and dy the variation of
the layer of the fluid
When evaporation takes place within an enclosed space, the partial pressure created by
the vapor molecules is called vapor pressure. Vapor pressure depend upon temperature
and increase with it.
The surface tension is the property of the liquid by virtue of which the free surface of
liquid at rest tends to have minimum area and as such it behaves as if covered with a
stretched membrane.
Therefore, surface tension of a liquid is measured as the force acting on unit length of
a line imagined to be drawn tangentially anywhere on the free surface of the liquid at
rest. It acts at right angles to this line on both the sides and a long tangent to the liquid
surface.
F
σ= (6.3)
L
The surface energy is defined as the amount of work done against the force of surface
tension in forming the liquid surface of a given area at a constant temperature.
W = σ · ∆A (6.4)
(f ) Capillarity
Rise or fall of liquid in a capillary tube (or in porous media) is caused by surface
tension and depends on the relative magnitudes of the cohesion of the liquid and the
adhesion of the liquid to the walls of the containing vessel. Liquids rise in tubes they wet
(adhesion > cohesion) and fall in tubes they do not wet (adhesion < cohesion).
The capillary rise (or depression) in a tube of the radius r is given approximately by
2σ cos θ
h= (6.5)
ωr
85
86 Lecture Notes of Physics for Engineers I Module
The bulk modulus of elasticity expresses the impressibility of a fluid. It is the ratio of
change in unit pressure to the corresponding volume change per unit of volume.
dp
K= (6.6)
−dV /V
Figure 6.5: The force of the fluid on the walls of the container is perpendicular to the
walls at all points.
The pressure at a specific point in a fluid can be measured with the device shown in
Figure 6.6.
The device consists of an evacuated cylinder enclosing a light piston connected to a spring.
The fluid pressure can be measured directly if the spring is calibrated in advance.
If the normal force exerted by the fluid is ∆F over a surface area of ∆A(look at the
Figure 6.7), then the pressure at that point is:
∆F dF
pB = lim = (6.8)
∆A→0 ∆A dA
Since pressure is force per unit area, it has unit of N · m−1 in SI system. Another name
for the SI unit of pressure is pascal ( Pa ). with 1 P a = 1 N · m2
To see this, let us select a portion of the fluid contained within an imaginary cylinder of
cross-sectional area A and height dy .
The upward force on the bottom of the cylinder is PA , and the downward force on the
top is (p + dp)A. The weight of the cylinder, the volume of which is dV , is given by
dW = ρgAdy, where ρ is the density of the fluid. Since the cylinder is in equilibrium, the
forces must add to zero, and so we get:
X dp
F = pA − (p + dp)A − ρgAdy = 0 ⇒ = −ρg (6.9)
dy
From this result, we see that an increase in elevation (positive dy) corresponds to a
decrease in pressure (negative dP ).
If P 1 and P 2 are pressures at the elevations y1 and y2 above the reference level (Figure 6.9),
and if the density is uniform, then integrating Equation 6.9 gives:
p1 − p2 = −ρg(y2 − y1 ) (6.10)
Figure 6.9: Pressure at the depth h below the free surface of a liquid
If the vessel is open at the top (see Figure 1.5.8), then the pressure at the depth h can be
obtained from Equation 6.10.
Taking atmospheric pressure to be pat = p2 , and noting that the depth h = y2 − y1 , we
find that the pressure below the free surface of a liquid is:
. An important application of Pascal’s law is the hydraulic press illustrated in Figure 6.10.
A force F1 is applied to a small piston of area A1 . The pressure is then transmitted
through a fluid to a larger piston of area A2 .
87
88 Lecture Notes of Physics for Engineers I Module
Figure 6.10: One of applications of Pascal’s law (a) Diagram of a hydraulic press. (b) A
vehicle undergoing repair is supported by a hydraulic lift in a garage.
F1 F2
Since the pressure is the same on both sides, we see that p = A1
= A2
. Therefore, the
force F2 is larger that F1 by the multiplying factor A
A1
2
.
F1 F2
= (6.12)
A1 A2
Figure 6.11: Measuring devices of pressure (a) Mercury manometer; (b) Mercury barometer
The pressure at point B equals pat = ρgh, where is the density of the fluid.
By the way, the pressure at B equals the pressure at point A, which is also the unknown
pressure P. Therefore, we conclude that:
Figure 6.12: (a) A totally submerged object that is less dense than the fluid in which it
is submerged experiences a net upward force and rises to the surface after it is released.
(b) A totally submerged object that is denser than the fluid experiences a net downward
force and sinks (c) An object floating on the surface of a fluid experiences two forces, the
gravitational force and the buoyant force.
Water provides partial support to any object placed in it. We say that an object placed
in a fluid is buoyed up by the fluid, and we call that upward force the buoyant force.
According to Archimedes’ principle, “the magnitude of the buoyant force always
equals the weight of the fluid displaced by the object”:
⃗ = |F⃗g |
|B| (6.14)
The buoyant force acts vertically upward through what was the center of gravity of the
displaced fluid (Figure 6.12).
Now imagine that the cube of water is replaced by a cube of steel of the same dimensions.
What is the buoyant force on the steel? The buoyant force acting on the steel is the same
as the buoyant force acting on the cube of water of the same dimensions. This result
applies for an immersed object of any shape, size, or density.
Let us show explicitly that the buoyant force is equal in magnitude to the weight of the
displaced fluid. The pressure at the bottom of the cube in Figure 1.5.11 is greater than
the pressure at the top by an amount ρf gh, where ρf is the density of the fluid and h is
the height of the cube.
Since the pressure difference, ∆p, is equal to the buoyant force per unit area, that is,
∆p = B A
, we see that B = (∆p)A = ρf gV , where V is the volume of the cubic volume
equivalent to the displaced liquid.
Since the mass of the fluid in the cube is M = ρf V , we see that:
B = Fg = ρf V g = M g (6.15)
where W is the weight of the displaced fluid.
It is now instructive to compare the forces on a totally immersed object with those acting
on a floating object.
Case I: A Totally Immersed Object
Where an object is totally immersed in a fluid of density ρf , the upward buoyant force is
given by B = ρf Vo g is the volume of the object. If the object has a density ρo , its weight
is equal to Fg = M g = ρo Vo g and the net force on it is B − Fg = (ρf − ρo )Vo g.
Hence, if the density of the object is less than the density of the fluid, the unsupported
object will accelerate upward (Figure 6.12a). If the density of the object is greater than
the density of the fluid, the unsupported object will sink (Figure 6.12b).
Case II: A Floating Object
89
90 Lecture Notes of Physics for Engineers I Module
Now consider an object in static equilibrium floating on a fluid, that is, one which is
partially immersed, as illustrated in the small figure below.
In this case, the upward buoyant force is balanced by the downward weight of the object.
If V is the volume of the fluid displaced by the object (which corresponds to that volume
of the object which is beneath the fluid level), then the buoyant force has a magnitude
given by B = ρf V g. Since the weight of the object is Fg = M g = ρo Vo g and Fg = B, we
see that ρf V g = ρo Vo g, or
ρo V
= (6.16)
ρf Vo
When fluid is in motion, its flow can be characterized as being one of two main types.
The flow is said to be steady if each particle of the fluid follows a smooth path, and the
paths of each particle do not cross each other. Thus, in a steady flow, the velocity of the
fluid at any point remains constant in time. Above a certain critical speed, fluid flow
becomes non-steady or turbulent. Turbulent flow is an irregular flow characterized
by small whirlpool-like regions. As an example, the flow of water in a stream becomes
turbulent in regions where rocks and other obstructions are encountered, often forming
“white water” rapids.
The term viscosity is commonly used in fluid flow to characterize the degree of internal
friction in the fluid. This internal friction is associated with the resistance to two adjacent
layers of the fluid to move relative to each other. Because of viscosity, part of kinetic
energy of a fluid is converted to thermal energy.
Because the motion of a real fluid is very complex, and not yet fully understood, we shall
make some assumptions in our approach. Many features of real fluids in motion can be
understood by considering the behavior of an ideal fluid. In our model of an ideal fluid,
we make the following four assumptions:
(b) Steady flow: In a steady flow, we assume that the velocity of the fluid at each point
remains constant in time;
(c) Incompressible fluid: The density of the fluid is assumed to remain constant in
time;
(d) Irrotational flow: Fluid is irrotational if there is no angular momentum of the fluid
about any point. If a small wheel placed anywhere in the fluid does not rotate about
its center of mass, the flow is irrotational. If the wheel were to rotate, as it would if
turbulence were present, the fluid would be rotational.
Consider now a fluid flowing through a pipe of non uniform size as in Figure 6.14. The
particles in the fluid move along the streamlines in steady flow.
In a small time interval ∆t, the fluid at the bottom end the pipe moves a distance
∆x1 = v1 ∆t. If M1 is the cross-sectional area in the region, then the mass contained in
the shaded region is ∆m1 = ρ1 A1 ∆x1 = ρ1 A1 v1 ∆t1
Figure 6.14: Steady flow motion of an incompressible fluid through a pipe of varying
cross-sectional area
Similarly, the fluid that moves through the upper end of the pipe in the time ∆t has a
mass ∆m2 = ρ2 A2 v2 ∆t2 . However, since mass is conserved and because the flow is steady,
the mass that crosses A1 in a time ∆t must equal the mass that crosses A2 in time ∆t.
Therefore ∆m1 = ∆m2 , or
ρ1 A1 v1 = ρ2 A2 v2 (6.17)
This expression is called the equation of continuity .
Since ρ is constant for an incompressible fluid, Equation 6.17 reduces to:
A1 v1 = A2 v2 (6.18)
That is, “the product of the area and the fluid speed at all points along the
pipe is a constant”. The product Av , which has the dimensions of volume
time
, is called
the volume flux, or flow rate.
91
92 Lecture Notes of Physics for Engineers I Module
W = (p1 − p2 )∆V
Part of this work goes into changing the kinetic energy of the fluid, and another part goes
into changing the gravitational potential energy. If ∆m is the mass passing through the
pipe in the time ∆t, then the change in its kinetic energy is:
1 1
∆K = (∆m)v22 − (∆m)v12
2 2
The change in its potential energy is:
∆U = ∆mgy2 − ∆mgy1
We can apply the work-energy theorem in the form W = ∆K + ∆U to this volume of
fluid to give:
1 2 1 2
(p1 − p2 )∆V = (∆m)v2 − (∆m)v1 + [∆mgy2 − ∆mgy1 ]
2 2
∆m
If we divide each term by ∆V , and recall that ρ = ∆V
, the previuos expression reduces to:
1 1
(p1 − p2 ) = ρv22 − ρv12 + ρgy2 − ρgy1
2 2
Rearranging terms, we get:
1 1
p1 + ρv12 + ρgy1 = p2 + ρv22 + ρgy2 (6.19)
2 2
The Equation 6.19 is the Bernoulli’s equation as applied to a nonviscous, incompressible
fluid in steady flow. It is often expressed as
1
p + ρv 2 + ρgy = Const. (6.20)
2
”Bernoulli’s equation says that the sum of pressure, p, the kinetic energy
per unit volume, 12 ρv 2 , and the potential energy per unit volume, ρgy, has the
same value at all point along streamline”
r2
v = v(r) = vo 1 − 2 (6.21)
R
where vo is the velocity at the center of the pipe.
Figure 6.16: Variation in velocity from the center to the linear walls of a fluid flowing
through a cylindrical pipe of radius R
In terms of the pressure difference ∆p = p1 − p2 across the length L of the pipe, the
central velocity is given by
∆pR2
vo = (6.22)
4ηL
By considering the flow through each cylindrical shell, it is possible to show that the total
mass flux dm
dt
(fluid mass flowing through the pipe per unit time) is:
dm ρπR4 ∆p πR4
= ⇒ D= ∆p (6.23)
dt 8ηL 8ηL
Consider a spherical object of radius r which undergoes a freely falling motion through a
viscous medium characterized by the coefficient of viscosity η, as illustrated in Figure 6.17.
If the speed v of the object is not large, the resistance force Fv to the object’s motion is
directly proportional to the speed and given by the following Stokes’ law :
Fv = 6πηrv (6.24)
In that motion, the force of viscosity Fv opposes its action to the weight W = mg of the
object. Since W > Fv the motion is downward and the object remains accelerated. When
Fv = W , the net force acting on the object becomes zero and that object will move with
a constant speed v L called the limit speed (example of some forms of rain).
93
94 Lecture Notes of Physics for Engineers I Module
It is the maximum constant velocity acquired by the body which falling freely in a viscous
medium. When a small spherical body falls freely through a viscous medium, three forces
act on it.
• Weight, W = 43 πr3 ρo g
2r2 (ρo − ρf )g
vt = (6.25)
9η
The critical velocity is that velocity of liquid flow, up to which the flow is streamlined,
and above to which the flow becomes turbulent.
The critical velocity is given by:
Kη
vc = (6.26)
ρr
with K: a dimensionless constant, r: the radius of the tube, ρ: the density of the liquid,
η: the coefficient of viscosity of the liquid.
According to Reynolds, the critical velocity is given by:
ηRe
vc = (6.27)
ρD
95
Chapter 7
OSCILLATORY MOTION
The vibration of a quartz crystal of modern electric watch, the swinging pendulum motion
of grandfather watch, the sound vibrations produced by a clarinet or any organ pipe, the
back-forth motion of the pistons in a car engine are all examples of repeated motions over
and over, called ”periodic motions” or ”oscillations”.
A period motion always has a stable equilibrium position. When a body in periodic
motion is moved away from equilibrium and then released, a force F⃗ or torque ⃗τ comes
into action to pull it back toward equilibrium. Meanwhile it picks up some kinetic energy,
so it overshoots, stopping somewhere on the other side, and again pulled back.
The spring vector force F⃗ responsible of the oscillatory motion is called ”a restoring force”.
an example is the motion of a simple pendulum or the motion of an elastic pendulum.
• The spring vector force F⃗ is the only horizontal net force acting on the body ; the
vertical normal reaction force and the gravitational force being balanced, adding to
zero vector.
• The coordinate system is defined so that the origin O is at the equilibrium position
(the spring being neither stretched nor compressed).
• If the body is displaced to the right to x = +A and then released, the net vector
force F⃗ and the acceleration vector a point to the left; the velocity v increases as
the body approaches to the equilibrium position O where the force F⃗ becomes zero
vector. But because of its motion, it overshoots the equilibrium position.
• On the other side of equilibrium position, the velocity vector v points to the left but
the acceleration vector a and the net vector force F⃗ point to the right; the velocity
v decreases until the body stops at position x = −A before coming back; etc...
96
Section 7.2: Simple Harmonic Motion 97
d2 x k d2 x
2
+ x=0 ⇔ 2
+ ω2x = 0 (7.1)
dt m dt
with
k
ω2 =
m
A solution of Equation 7.1 is:
(a) Definitions
By definition, a particle moving along x−axis has a Simple Harmonic Motion when the
particle’s displacement x relative to the origin of the coordinate system is given by the
relation
97
98 Lecture Notes of Physics for Engineers I Module
• Xm or A is the motion’s amplitude, that means the maximum value that can be
taken by the abscissa x can take, in meters.
dx dx
v= = ωXm cos(ωt + φ); or v = ωA cos(ωt + φ) (7.4)
dt dt
The instantaneous acceleration, a, of a particle in a simple harmonic motion is given
by:
d2 x
a= 2
= −ω 2 Xm sin(ωt + φ) = −ω 2 A sin(ωt + φ) ⇔ a = −ω 2 x (7.5)
dt
The relationship between instantaneous velocity v, instantaneous position x and the
amplitude Xm can be given by:
x2 v2
2
= sin2 (ωt + φ) and = cos2 (ωt + φ) ⇔
Xm ω 2 Xm
2
x2 v2 x2 v2
2
+ = sin2 (ωt + φ) + cos2 (ωt + φ) ⇔ + =1 ⇔
Xm ω 2 Xm
2 2
Xm ω 2 Xm
2
ω 2 x2 + v 2 ω 2 Xm
2
= ⇔ ω 2 x2 + v 2 = ω 2 X m
2
ω 2 Xm2 ω 2 Xm
2
p
⇒ v = ±ω 2 − x2
Xm (7.6)
In addition, the amplitude Xm and the phase angle φ can be determined by initial
conditions, at time t = 0:
x20 v02
2
= sin2 φ and 2 2
= cos2 φ ⇔
Xm ω Xm
x20 v02 x20 v02
2
+ 2 2 = sin2 φ + cos2 φ ⇔ 2
+ 2 2 =1 ⇔
Xm ω X m Xm ω Xm
ω 2 x20 + v02 ω 2 Xm
2
= 2 2 ⇔ ω 2 x20 + v02 = ω 2 Xm
2
ω 2 Xm 2 ω Xm
r
v02
⇒ Xm = x20 + (7.7)
ω2
But also:
x0 v0 ωx0
= sin φ ⇔ = cos φ ⇔ tan φ =
Xm ωXm v0
−1 ωx0
⇒ tan (7.8)
v0
The simple harmonic motion (SHM) is always a periodic motion, with a time period
T and a frequency υ, defined by:
2π 1 2π
T = = ⇔ ω= = 2πν = 2πf (7.9)
ω ν T
☛ The time period T is the duration of one complete oscillation, after what the
particle has the same position and the same algebraic value of abscissa x . The SI
unit of the period T is the second (1 s).
☛ The frequency f = ν (also called the number of cycles or periods per second) is
expressed in hertz (1 Hz).
n
X
F⃗i = F⃗ = m⃗a ⇒ F⃗ = m⃗a and ⃗a = −ω 2⃗x ⇒ F⃗ = −mω 2⃗x = −k⃗x (7.10)
i=1
☛ The net vector force F⃗ = −k⃗x occurs when a spring (elastic body) is deformed by
an external force; it is then called the restoring force of Hooke.
☛ Then, the time period T and the frequency ν of a simple harmonic motion can be
related to the strength k of a spring by:
r r
m 1 k
T = 2π and ν= = (7.12)
k 2π m
99
100 Lecture Notes of Physics for Engineers I Module
☛ Kinetic energy K will be maximum, when the particle passes at the equilibrium
position where x = 0 and v = vm .
We know that for a force F⃗ that derives from a potential (a conservative force)
dU
U =− ; with F = −kx ⇔ dU = kxdx
dx
The potential energy U of a particle in a simple harmonic motion becomes:
Z U Z x
1 1
U= dU = kxdx ⇒ U = kx2 = mω 2 x2 (7.14)
0 0 2 2
By definition, the total mechanical energy E of a particle equals to the sum of its
kinetic energy and potential energy.
1 1 1
E = K + U = mω 2 Xm 2
− x2 + mω 2 x2 = mω 2 Xm 2
= Constant (7.15)
2 2 2
Since m, ω and Xm are constant, the total mechanical energy E of a particle in a simple
harmonic motion is also constant; we say that E is conserved .
The mechanical energy is constant (as it is required by the law of conservation of energy).
The kinetic energy and potential energies are shown as functions of time in Figure 7.3a
and as functions of displacement in Figure 7.3b.
Note that the total energy E depends on the amplitude Xm which, in turn, depends on
the initial conditions (Equation 7.7).
The total energy of the Simple Harmonic Oscillator (mass+spring) depends only on the
spring constant k and the amplitude A. The total energy (mechanical) of a linear oscillator
is constant and time–independent.
Figure 7.4: Relationship between the uniform circular motion of a point P and the simple
harmonic motion of a point Q. A particle at P moves in a circle of radius A with constant
angular speed ω.
−−→
The projection Q of P onto OX executes a simple harmonic motion, and the x−
coordinate of the particle P at any time t is:
x = A cos θ (7.16)
Knowing that in a uniform circular motion, the constant radial acceleration vector ⃗ar = ⃗a
always points toward the origin O, a = −ω 2 A and that the x –component of ⃗a is the
−−→
acceleration of the SHM of Q along OX given by:
101
102 Lecture Notes of Physics for Engineers I Module
x
F = −mg sin( ) (7.20)
l
d2 x x
m 2
= −mg( ) (7.23)
dt l
or
d2 x x
2
+ g( ) = 0 (7.24)
dt l
Therefore, for small oscillations, the position x along the arc of the path is given by
r
g
x(t) = A sin t+ϕ
l
2π
Since T = ω
, we have
s
l
T = 2π (7.25)
g
and
r
1 1 g
f= = (7.26)
T 2π l
For a simple pendulum swinging at small displacement x, Xmax /l, in terms of the angle
that the string makes with the vertical, the maximum angle θmax should be less that
0.1radians.
τnet = Iα (7.27)
Here, the quantity relating the angular acceleration and the torque is the moment of inertia
I of mass. The moment of inertia is not always a scalar quantity, thus Equation 7.27 is
not always a linear relation. The moment of inertia is chosen with respect to any axis of
rotation.
d2 θ
I = −mgl sin θ (7.29)
dt2
2
where α = ddt2θ
For the small angles, Equation 7.29 becomes the differential equation of the oscillating
system:
d2 θ mglθ
= − (7.30)
dt2 I
The angular frequency ω and the period T are, respectively
r
mgl
ω= (7.31)
I
and
s
I
T = 2π (7.32)
mgl
Which is exactly Equation 7.25 ot the expression od the period of a simple pendulum.
103
104 Lecture Notes of Physics for Engineers I Module
τ = −κθ (7.33)
Here κ > 0 is a constant called the torsion constant, which depends on the length,
diameter, and material of the suspension wire. The negative sign indicates that the
torque acts in the opposite direction of the angle of the torsion. The above
equation is the angular form of Hooke’s law. The rotational equation of motion os the
system is written as
d2 θ
I =τ (7.34)
dt2
Where I is the moment of inertia of the disk. The moment of inertia of the wire is assumed
to be negligible. Combining the two previous equations, we have
d2 θ
I 2 + κθ = 0 (7.35)
dt
Equation 7.30 is clearly a simple harmonic motion equation. Hence, we can write the
standard solution as
θ = A cos(ωt + φ0 )
where
r
κ
ω= (7.36)
I
The period of the angular simple harmonic oscillator is
r
I
T = 2π (7.37)
κ
The torsion pendulum is often used for time-keeping purposes. For instance, the balance
wheel in a mechanical wristwatch is a torsion pendulum in which the restoring torque is
provided by a coiled spring.
m = ρAL (7.38)
Figure 7.8 shows the instantaneous position of the mass of water in the U–tube. The
displacement of the water from its equilibrium level is represented by x measured in one
of the arms of the U–tube. The restoring force acting on the whole mass of water is then
just the weight of the displaced water in the right hand arm of column height 2x:
F = −ρ2xAg (7.39)
The second Newton’s law can be written as
d2 x
ρAL = −ρ2xAg (7.40)
dt2
Note that all the mass is in motion. This can be written in the usual form of SHM
equation.
d2 x 2g
+ x=0 (7.41)
dt2 L
Which defines the angular frequency of SHM
r s
2g l
ω= or T = 2π (7.42)
L 2g
105
106 Lecture Notes of Physics for Engineers I Module
b = Kη
☛ K is a coefficient characterizing the shape and size of a body moving in a viscous
fluid. (e.g.: K = 6πR for a spherical body of radius R).
☛ η is a coefficient characterizing the fluid’s nature, and called ”coefficient of
viscosity” or ”stickiness”.
Note that Equation 7.43 is only verified in the case of a laminar flow (regular) of the
fluid around a solid body.
➯ When the flow is turbulent (formation of whirlpools behind the solid body), the
force of viscous friction F⃗v increases quickly in function of the body’s velocity ⃗v and
its magnitude becomes:
F v = b1 v + b2 v 2 + b3 v 3 + · · · (7.44)
➋ the restoring elastic force F⃗ = F⃗0 + k⃗x exerted on the body by the spring, and with
magnitude F = Fo − kx;
Applying the Newton’s Second law for Dynamics of translation and considering that the
Archimedes’ force is negligible, we get
n
X
F⃗i = m⃗a ⇒ F⃗ + P⃗ + F⃗v = m⃗a
i=1
F − P − Fv = ma; F = F0 − kx and F0 = P = mg
at equilibrium
mg − kx − mg − bv = ma ⇒ ⇔ −kx − bv = ma
dx d2 x
where v = dt
and a = dt2
dx d2 x
−kx − b
=m 2
dt dt
The differential equation of the damped oscillations then becomes:
d2 x b dx k
2
+ + x=0 (7.45)
dt m dt m
But for more simplicity, we use
b k
and ω02 =
2δ = (7.46)
m m
Putting Equation 7.46 into Equation 7.45, the differential equation of damped
oscillations can also be expressed by
d2 x dx
2
+ 2δ + ω02 x = 0 (7.47)
dt dt
where δ is the damping factor and ω0 is the angular frequency of a simple
harmonic motion associated to the damped oscillations.
Equation 7.45 and Equation 7.47 of damped oscillations, differ to Equation 7.1 of a simple
harmonic motion only by the term in dx
dt
, it is the reason why we call it the differential
equation of a damped oscillator (m–k–b).
A solution of equations (7.45) and (7.47) is
107
108 Lecture Notes of Physics for Engineers I Module
However, the amplitude Xm (t) is here a decreasing function of time t, whereas the
phase angle φ of damped oscillations depends only on the motion’s initial conditions
t = 0.
The amplitude Xm (t) of damped oscillations decreases exponentially in function of time
according to the following equation:
❖ C and ω must be related to the constants of the m–k–b system, whereas Xm0 and
φ depend more on the initial abscissa x0 and velocity v0 (t = 0).
2
❖ Putting expressions of x, dx
dt
and ddt2x from Equation 7.49 into Equation 7.48, we obtain
the conditions in which the differential equation of damped oscillations is
satisfied.
dx
= −Xm0 ωe−Ct sin(ωt + φ) − Xm0 Ce−Ct sin(ωt + φ) (7.50)
dt
and
d2 x −Ct
= X m0 −ω 2
+ C 2
e cos(ωt + φ) + 2Xm0 ωCe−Ct sin(ωt + φ) (7.51)
dt2
Putting equations (7.50), (7.51) and (7.48) into Equation 7.47, we get
−ω 2 + C 2 − 2δC + ω02 = 0
and
2ωC − 2δω = 0
b
⇔C=δ=
2m
⇒ ω 2 = ω02 − δ 2 (7.53)
➡ The angular frequency ω and the frequency ν of damped oscillations are respectively
given by:
s 2
δ
q
2 2
ω = ω0 − δ = ω0 1 − (7.54)
ω0
and
s 2 s 2
ω ω0 δ δ
ν= = 1− = ν0 1− (7.55)
2π 2π ω0 ω0
➡ The angular frequency q ω of damped oscillations is always less than the
k
angular frequency ω0 = m of a simple harmonic motion (non damped oscillations). It
will be all the smaller if the damping factor δ is all the bigger.
➡ The decreasing of the amplitude Xm (t) od damped oscillations per period T is mea-
sured by the fraction XXmm0
(T )
= e−δT called the logarithmic decrement of damped
oscillations; this fraction is always comprised between 0 and 1.
➡ The period T of damped oscillations is called the pseudo–period .
Figure 7.11: Decreasing of the pseudo-period, with respect to the increasing of the damping
factor
✎ If δ < ω0 , the angular frequency ω of damped oscillations decreases when the damping
factor δ increases (Figure 7.11 a–b–c–d), we observe underdamping oscillations.
This is applied in balances, needles of any electrical laboratory apparatus like an
ammeter or a voltmeter.
➨ Let’s x′ = Xme cos(ωe t) be the simple harmonic motion equation of the spring’s
suspension point P . The amplitude of the driving motion is Xme is constant, but its
driving angular frequency ωe varies.
109
110 Lecture Notes of Physics for Engineers I Module
✎ If the driving angular frequency ωe is too small (ωe << ω0 ), the strength k of the
spring is such as the displacement x of mass m, will be quite the same as the
displacement x′ of the suspension point P .
✎ If ωe >> ω0 , the displacement x of mass m becomes too small, due to its inertia.
Figure 7.13: Variation of the responding amplitude, in function of the driving angular
frequency ωe and for three damping factors
dx
Fv = −bv = −b
dt
Using the Newton’s Second law, we have:
dx dx d2 x
−b − k(x − x′ ) = ma ⇒ −b − kx + kx′ = m 2 ;
dt dt dt
and also
d2 x b dx k k
2
+ + x = x′
dt m dt m m
b k 2 ′
with m = 2δ; m = ωo and x = Xme cos(ωe t)
The differential equation of forced or driven oscillations becomes
d2 x dx
+ 2δ + ωo2 x = ωo2 Xme cos(ωe t) (7.56)
dt2 dt
Equation 7.56 has the same shape like Equation 7.47, but its second member is different
to zero. Such an equation is called a nonhomogenous differential equation; and its
solution is given by:
x = Xm cos(ωe t + θ) (7.57)
Where Xm and θ , are the amplitude and the phase angle (initial angular phase) of forced
or driven oscillations, respectively.
Putting expressions of the velocity v = dxdt
= −Xme ωe sin(ωe t + θ) and of the acceleration
d2 x 2
v = dt2 = −Xme ωe cos(ωe t + θ) in Equation 7.56, we find the expressions of Xm and θ of
the forced or driven oscillations, as functions of the driving angular frequency ωe .
Note
➊ The amplitude Xm of force oscillations takes a maximum value XM when ωe = ωm and
quite less that ω0 ; so that for Xm = XM , we get
s 2
δ
ωm = ω0 1 − 2 (7.60)
ω0
✎ There will be no resonance (ωm = 0), when
δ 1
≥√
ω0 2
✎ If the damping δ is very small, the ωδ0 ≃ 0 and ωm ≃ ω0
➋ We call the fact of quality of forced oscillator (m − k − b), the physical quantity:
ω0 ω0 ω0 m
Q= ≃ = (7.61)
ωe 2δ b
111
112 Lecture Notes of Physics for Engineers I Module
✎ If b is small, then Q is high and the system will be more resonant; we say that the
resonance is sharp. The resonance will be flat in the opposite case.
8.1 Introduction
In this chapter we will only study the propagation of waves in a deformable (or elastic)
medium; that kind of waves are called mechanical waves.
To produce mechanical waves, we can displace a section of the propagation medium from
its position of equilibrium. Then, cause of the elastic properties of that medium, that
section starts to vibrate around the equilibrium position.
The disturbance (deformation) produced travels through the medium, from one section to
the closer one, putting it in the same vibration but at successively later time. We say
that there is vibration propagation or simply a progressive wave.
➢ A wave is longitudinal when on its passage, the particles of the propagation medium
oscillate in the same direction as the wave direction of propagation. An example is the
propagation of a compression–extension along a stretched spring.
According to the number of space dimensions in which the wave propagates, waves can be:
113
114 Lecture Notes of Physics for Engineers I Module
Note
❶ When a wave propagates in a medium, the particles of that medium oscillate at the
same place, without moving in the wave’s direction.
❹ The propagation velocity ⃗v of a wave is called the wave speed or the wave
celerity, and it only depends on the properties of the propagation medium.
❺ It is the elasticity of the propagation medium that produces in each point, a restoring
elastic force F⃗ = −k⃗x that tend to restore each medium’s section to its equilibrium
position.
Figure 8.4: Photos of a same harmonic wave along a stretched rope, taken at different
times
Let at time t0 = 0, the physical quantity associated to the wave propagation be described
by Ψ(x, 0) = Ψ(x) in the reference frame O. Imagine also another reference frame O′ that
has the same motion as the wave, but that occupies the position O at time t0 = 0 (e.g.: a
deformation of a vibrating rope, a compression of a spring). The nature of the wave
function Ψ(x, t) only depends on the kind of wave (mechanical wave along a string
or a spring, mechanical wave at the surface of a liquid, sound wave in air, electromagnetic
or light wave in a medium).
At any later time t, that physical quantity is characterized by Ψ(x′ ) = Ψ(x − d). And since
the two quantities describe a same wave, but delayed in time by t; the one–dimension
−−→
wave function for a wave traveling in the positive direction of OX is:
Note
✎ For a wave traveling in the positive direction, we use the negative (−) sign, but for a
wave traveling in the negative direction we use the positive (+) sign.
✏ At a fixed time t (like in the case of taking a snapshot of the pulse), the wave function
becomes Ψ(x) ; it is called the waveform and it defines a curve representing the geometric
shape of the pulse at that time.
Figure 8.5: One–dimension harmonic wave traveling to the right at a wave velocity v
A harmonic or sinusoidal wave is defined as a wave that results from the propagation
of a simple harmonic motion in a given medium.
A harmonic wave is characterized by the following physical quantities:
➊ The wave amplitude: the maximum value that can take the physical quantity
associated to the wave.
➌ The wave period T or temporal period: the time duration of one complete
vibration; it is expressed in seconds (1 s).
➋ The wavelength λ or spatial period: the distance covered by the wave during one
period of time T ; it is expressed in meters (1 m).
➍ The wave frequency ν: the number of vibrations or cycles per second; it is expressed
in hertz (1 Hz).
115
116 Lecture Notes of Physics for Engineers I Module
Ψ(x, t) = Ψ(x ∓ vt) = Ψm sin [k (x ∓ vt)] or Ψ(x, t) = Ψ(x ∓ vt) = Ψm cos [k (x ∓ vt)]
(8.4)
where Ψm is the harmonic wave amplitude: the maximum value that can take the
physical quantity Ψ(x, t) associated with the wave.
Since the sine or cosine function is a periodic function of period 2π, the harmonic
wave function Ψ(x, t) that propagates, takes again the same algebraic value and the same
variation, when the wave has traveled a distance equal to the wavelength λ ; this occurs
after a time that equals to one period T .
The motion of an element mass ∆m of the stretched string, reached by the wave front
travelling and having an element length ∆l, is composed by two perpendicular motions:
−−→
➊ A uniformly varying rectilinear motion along the OY axis, with acceleration ⃗ay :
Ty T sin θ T sin θ
ay = = and vy = ay ∆t ⇒ vy = × ∆t (8.11)
∆m ∆m ∆m
−−→
➋ A uniformly varying rectilinear motion along the OX axis, such as:
∆l ∆l
∆l = vx ∆t ⇔ ∆t = = (8.12)
vx v
Putting Equation 8.12 in Equation 8.11, where vx = v, the wave celerity along the string,
we get:
T sin θ ∆l ∆l
vy = × with vy = v tan θ ⇒ v 2 = vx2 = T × × cos θ
∆m v v
The wave celerity v along a stretched string or rope is given by the following general
expression:
s
T
v= cos θ (8.13)
µ
However, the angle θ is generally very small; then cos θ ≈ 1
The wave celerity v along a stretched rope or string simply becomes:
s
T
v= (8.14)
µ
Note that the mechanical wave celerity always depends on elastic and inertial
properties of the medium of propagation.
117
118 Lecture Notes of Physics for Engineers I Module
Figure 8.8: Reflection of a wave pulse by non deformable and deformable obstacles
8.6 Refraction of a Wave Pulse
At the boundary of two media with different inertia, an incident wave pulse sets up:
✎ Weak refracted wave pulse, vibrating in the same direction as the incident wave pulse;
and if the second medium’s inertia is the highest: v2 < v1 .
✎ A weak refracted wave pulse and a strong reflected wave pulse, both vibrating in
the same direction as the incident wave pulse; and if the second medium’s inertia is the
smallest: v2 > v1 .
E εVo 1 1
P = = = εvS = ρω 2 vSΨ2m = (ρω 2 vS)A2 (8.17)
t t 2 2
where:
➡ Vo = Sd = Svt,
∂Ψ ∂Ψ ∂u ∂Ψ ∂ 2Ψ ∂ 2Ψ
= = and = (8.21)
∂x ∂u ∂x ∂u ∂x2 ∂u2
then
∂Ψ ∂Ψ ∂u ∂Ψ ∂ 2Ψ 2
2 ∂ Ψ ∂u
2
2∂ Ψ
= = ∓v and = (∓v) = v (8.22)
∂t ∂u ∂t ∂u ∂t2 ∂u2 ∂t ∂u2
that is
∂ 2Ψ 1 ∂ 2Ψ
= (8.23)
∂u2 v 2 ∂t2
Comparing Equation 8.22 and Equation 8.23, we get:
∂ 2Ψ 1 ∂ 2Ψ
= (8.24)
∂x2 v 2 ∂t2
The equation (2.52) is the one-dimension wave differential equation of a wave
−−→
that travels along the OX axis.
Note
➊ In the particular case of light waves, a part of electromagnetic waves, of the wave length
λ in the ranges 0.4µm ≤ λ ≤ 0.7µm; the general wave function Ψ(x, t) is replaced by the
electric field E or the magnetic field B that compose any electromagnetic wave, travelling
−−→
in the OX direction.
119
120 Lecture Notes of Physics for Engineers I Module
∂ 2E 1 ∂ 2E ∂ 2B 1 ∂ 2B
= or = (8.25)
∂x2 v 2 ∂t2 ∂x2 v 2 ∂t2
where E = Em sin(kx − ωt); B = Bm sin(kx − ωt) and v –the celerity of light in that
medium, characterized by its refractive index n.
➋ For sound waves, the general wave function Ψ(x, t) is replaced by the pressure fluctuation
∆p or p(t) in the propagation medium. The one–dimension sound wave differential equation
−−→
along OX direction becomes:
∂ 2 ∆p 1 ∂ 2 ∆p ∂ 2p 1 ∂ 2p
= 2 or = 2 2 (8.26)
∂x2 v ∂t2 ∂x2 v ∂t
A compression (∆p > 0) corresponds to a minimum displacement of a section of the
medium where sound waves propagate; and an expansion (∆p < 0) corresponds to a
maximum displacement of a section of the medium where sound waves propagate.
➌ In general, the three–dimension wave differential equation becomes:
∂ 2Ψ ∂ 2Ψ ∂ 2Ψ 1 ∂ 2Ψ
+ + = (8.27)
∂x2 ∂y 2 ∂z 2 v 2 ∂t2
[1] Raymond A. Serway and John W Jewett, Jr, Physics for Scientists and Engineers
with Modern Physics, 2nd Edition, Brooks/Cole (2010).
[3] Frederick J, Bueche and Eugene Hecht, Schaum’s Easy Outlines of College Physics,
McGraw-Hill (2000)
[5] A. S. Kompaneyets, Pheoretical Physics, Dover Publications, Inc. Mineola, New York
(2012)
[6] Late Dr. Baziruwiha Jean de Dieu , Mechanics, Kigali Institute of Education (2013)
[7] Dr. Lakhan Lai Yadav, Dr. Evariste Minani and late Dr. Jean de Dieu Baziruwiha
(1964-2013), Kinetic Theory of Matter and Thermodynamics, Kigali Institute of
Education (2014)
[8] Richard Wolfoson, Essential University Physics ,2nd Edition, Addison-Wesley (2012)
121