Você está na página 1de 64

Project Title

STUDIES ON SOME BASIC CONCEPTS OF


MECHANICS

A project submitted to the Department of Mathematics


for fulfillment of the requirement for course no MAT 421
(Course title: Project in Mathematics and Relative fields)
in 4th Year 2nd Semester 2017 (Session 2013-14)
for B.Sc. (Honors) degree in Mathematics

Supervised By
Dr. Md Gulam Ali Hayder Chowdhury
Professor,
Dept. of Mathematics
SUST, Sylhet

Submitted By
Pritom Sikder, Reg. no.: 2013133072
Md. Shafiqul Islam, Reg. no.: 2013133079
Sohel Ahammed, Reg. no.: 2013133105

Department of Mathematics
Shahjalal University of Science and Technology, Sylhet-3114
Declaration

We the undersigned hereby declare that, we have done this project, entitled “Studies on Some
Basic Concepts of Mechanics” under the Supervision of Professor Dr. Md Gulam Ali Hayder
Chowdhury, Dept. of Mathematics, SUST.

___________________________________ __________________________________
Pritom Sikder Md. Shafiqul Islam
Reg. no: 2013133072 Reg. no: 2013133079

___________________________________
Sohel Ahammed
Reg. no: 2013133105
Certification

We hereby certify that this project, entitled ‘Studies On Some Basic Concepts Of Mechanics
(Moment Of Inertia, Center Of Gravity, Virtual Work, Equations Of Motion)’, submitted
by the students Pritom Sikder (Reg no. 2013133072), Md. Shafiqul Islam (Reg no. 2013133079),
Sohel Ahammed (Reg no. 2013133105) conforms to acceptable standards and is fully adequate
in scope and quality to fulfill the requirements for Course No. MAT 421 ( Course Title: Project
in Mathematics and Relevant Fields) in 4th year 2nd semester 2017 (session 2013-14) of
B.Sc.(Honors) degree in Mathematics.

Dr. Md Gulam Ali Hayder Chowdhury


Professor,
Department of Mathematics
SUST, Sylhet
ACKOWLDGEMENT

First of all we are grateful to The Almighty God for establishing us to complete this project.
The research project has been conducted under the constant supervision of Prof. Dr. M. G. A.
Hayder Chowdhury, Department of Mathematics, Shahjalal University of Science &
Technology, Sylhet.
We offer our profound love and gratitude to our beloved teacher, Prof. Dr. M. G. A. Hayder
Chowdhury for providing us with the necessary suggestions and guidelines for our research
project, “Studies on Some Basic Concepts of Mechanics”.
We are thankful to and fortunate enough to get constant encouragement, support and guidance
from all Teaching staffs of Department of Mathematics which helped us in successfully
completing our project work.
Last but not least we are indebted to our beloved parents and family members for their immense
love, affection and financial support which help us to complete our work.

March, 2019 Author

Pritom Sikder
Md. Shafiqul Islam
Sohel Ahammed
CONTENTS

Introduction

List Of Contents

CHAPTER ONE: MOMENT OF INERTIA


1.1 History………………………………………………………………………………………1
1.2 Definition…………………………………………………………………………………....1
1.3 Radius of gyration……………………………………………………………………….....1
1.4 Product of inertia…………………………………………………………………………...1
1.5 How does rotational inertia related to Newton’s 2nd law ?.................................................1
1.6 How long will the fly wheel take to reach a steady speed if starting from rest?..............2
1.7 How can we calculate moment of inertia in general?.........................................................3
1.8 How can we find the moment of inertia of complex shapes?.............................................4
1.9 Parallel axis theorem with related problems……………………………………………...5
1.10 Perpendicular axis theorem……………………………………………………………....8
1.11 Problems…………………………………………………………………………………..10

CHAPTER TWO: CENTER OF GRAVITY


2.1 Introduction………………………………………………………………………………...13
2.2 History……………………………………………………………………………………....13

2.3 Definition……………………………………………………………………………………14

2.4 Definitive Intuition…………………………………………………………………………14

2.5 Equilibrium Toys…………………………………………………………………………..14


2.6 General formula for the determination of the center of gravity………………………...15

2.7 General formula for the determination of the center of gravity by integration………..16

2.8 Formula for the center of gravity of an arc……………………………………………….16

2.9 Why does center of gravity important?...............................................................................18

2.10 Center of gravity of different dimensional shapes……………………………………....18

2.11 The difference between center of mass and center gravity……………………………..20

2.12 Problems…………………………………………………………………………………...22

CHAPTER THREE: VIRTUAL WORK


3.1 History………………………………………………………………………………………27
3.2 Work of a force……………………………………………………………………………..27
3.3 Work of a couple moment………………………………………………………………….28
3.4 Definition……………………………………………………………………………………29
3.5 Principle of virtual work…………………………………………………………………...30
3.6 Principle of virtual work for a system of connected rigid bodies………………………..31
3.7 Procedure for analysis……………………………………………………………………...32
3.8 D’Alembert’s principle……………………………………………………………………..33
3.9 problems……………………………………………………………………………………..34

CHAPTER FOUR: EQUATIONS OF MOTION


4.1 History………………………………………………………………………………………41
4.2 Related definitions………………………………………………………………………….43
4.3 Newton laws of motion with explanation…………………………………………………44
4.4 One dimensional motion…………………………………………………………………...48
4.4.1 uniform acceleration………………………………………………………………….48
4.4.2 constant translational acceleration in a straight line……………………………….48
4.4.3 Related problems…………………………………………………………………….49
4.5 Two dimensional motion………………………………………………………………….51
4.5.1 Projectile motion……………………………………………………………………52
4.5.2 Velocity, acceleration and displacement of the projectile………………………..53
4.5.3 problems…………………………………………………………………………….55

CHAPTER FIVE: CONCLUSION………………………………………………………….56


Introduction
Mechanics is a branch of the physical sciences concerned with the state of rest or motion of
bodies that are subjected to the action of forces. The study of mechanics involves many more
subject areas. However, initial study is usually split into two areas; statics and dynamics
Our project titled as “Studies on some basic concepts of Mechanics” is mainly based on the
Statics and Dynamics of Newtonian Mechanics. Here we have mainly focused on the Moment
of Inertia, Equations of motion. from the Dynamics. Other than that, Centre of Gravity, Virtual
Works form the Statics have also been highlighted here.
CHAPTER ONE: MOMENT OF INERTIA

1.1 History: In 1673 Christiaan Huygens introduced this parameter in his study of the
oscillation of a body hanging from a pivot, known as a compound pendulum. The term moment
of inertia was introduced by Leonhard Euler in his book Theoria motus corporum solidorum seu
rigidorum in 1765, and it is incorporated into Euler's second law.

1.2 Definition:

Moment of inertia of a body about a line –


If the mass of every particle of a material system be multiplied by the square of the perpendicular distance
from a straight line the sum of the products so formed is called the moment of inertia of the system that
line.

If m be the mass of an element of the system, r is the distance from the given line, the moment of
inertia about that line = ∑mr2. In-short we shall write M.I for moment of inertia.

1.3 Radius of Gyration:

If ∑mr2 = MK2 when M is the mass of the body, K is called radius of gyration or sometimes it is
called Swing Radius about the given line.

1.4 Product of Inertia:

Let (x, y) be the coordinate of a particle of mass M with respect to two fixed lines at right angles
as coordinate axis. Take the product x × y and the corresponding products for the other particles
of the body. The sum of the products i.e., ∑ x × y is called the product of inertia of the body
about this lines.

1.5 How does rotational inertia relate to Newton's 2ⁿᵈ law?

Rotational inertia takes the place of mass in the rotational version of Newton's 2ⁿᵈ law.
Consider a mass m attached to one end of a massless rod. The other end of the rod is hinged so
that the system can rotate about the central hinge point as shown in Figure 1.

1|Page
Figure 1: A mass rotating due to a tangential force.

We now start rotating the system by applying a tangential force FT to the mass. From Newton’s
2ⁿᵈ law,
FT = maT
this can also be written as
FT = m(rα)
Newton's 2ⁿᵈ law relates force to acceleration. In rotational mechanics torque τ takes the place of
force.

Problem:

A motor capable of producing a constant torque of 100 Nm and a maximum rotation speed
of 150 rad/s is connected to a flywheel with rotational inertia 0.1 kgm2. What angular
acceleration will the flywheel experience as the motor is switched on?
Solution:

Rearranging the rotational version of Newton's 2ⁿᵈ law and substituting in the numbers we find:
α = τ/I
=100 Nm/0.1 kgm2
=1000 rad/s2

1.6 How long will the flywheel take to reach a steady speed if starting from
rest?

Using rotational kinematics,


ω = 𝜔0 +αt
Since we know the maximum rotational velocity of the motor, we can solve to find the time
taken to accelerate up to that rotational velocity.

2|Page
t = ωmax/α
=(150 rad/s)/(1000 rad/s2)
=0.15 s

1.7 How can we calculate moment inertia in general?

Often mechanical systems are made of many masses connected together, or complex shapes.
It is possible to calculate the total rotational inertia for any shape about any axis by summing the
rotational inertia of each mass.

Figure 2: A rigid system of masses shown with two different rotation axes

Consider the object shown in figure 2(a). What is its rotational inertia?

Summing all the mr2 terms for each mass,

I = (1 kg⋅12 m2) + (1 kg⋅1.52 m2) + (1 kg⋅0.752 m2) + (2 kg⋅0.752 m2)

= 4.9375 kg⋅m2

Consider the alternate case of Figure 2(b) of the same system rotating about a
different axis. What would you expect the rotational inertia to be in this case?

Although the system of masses is the same as before, they are now rotating about a much closer
axis. Because of the dependence on the square of the distance to the rotational axis, we would
expect the rotational inertia to be significantly lower.

3|Page
I = (2 kg⋅0.52 m2)+(1 kg⋅0.52 m2)+(1 kg⋅0.52 m2)+(1 kg⋅0.52 m2)

= 1.25 kg⋅m2

1.8 How can we find the moment inertia of complex shapes?

For more complicated shapes, it is generally necessary to use calculus to find the rotational
inertia. However, for many common geometric shapes it is possible to find tables of equations
for the rotational inertia in textbooks or other sources. These typically give the moment of inertia
for a shape rotated about its centroid (which often corresponds with the shapes center of mass).
For example, the rotational inertia of a solid cylinder with radius r rotated about a central axis is
1
I = 2mr2

and for a hollow cylinder with inner and outer radii ri and r0 respectively,
𝑚( 𝑟𝑖2 +𝑟02 )
I= 2

Expressions for other simple shapes are shown in Figure 3.

Figure 3: Equations for the rotational inertia of some simple shapes under rotation.

4|Page
Complex shapes can often be represented as combinations of simple shapes for which there exists a
known equation for rotational inertia. We can then combine these rotational inertia to find that of the
composite object.
It is also sometimes possible to calculate rotational inertia for uniform density objects with voids by
considering the void to be a shape with negative rotational inertia.
The problem that we will likely run into when combining simple shapes is that the equations tell us the
rotational inertia as found about the centroid of the shape and this does not necessarily correspond to the
axis of rotation of our composite shape. We can account for this using the parallel axis theorem.

The parallel axis theorem allows us to find the moment of inertia of an object about a point o as long as
we known the moment of inertia of the shape around its centroid c, mass m and distance d between
points o and c.
Io = Ic+md2

1.9 Parallel axes theorem:

Purpose:

The parallel axis theorem is the theorem determines the moment of inertia of a rigid body about
any given axis, given that moment of inertia about the parallel axis through the center of mass of
an object and the perpendicular distance between the axes. The moment of inertia of any object
can be determined dynamically with the Parallel Axis Theorem.
Statement:
The product of inertia of body about two perpendicular axis is equal to its product of inertia
about parallel axis through its center of inertia, together with the product of inertia of the whole
mass, placed at the center of inertia, about the original axis.
Proof:
Let G be the center of gravity of the body and the three perpendicular axis through G be GX,
GY, GZ respectively. Let OX1, OY1, OZ1 be any three parallel axis through any point O. Let an
element m be placed at P. The coordinates of the element with respect to G be (x,y,z) and with
respect to O (x1, y1,z1). Let the coordinates of G with respect to O be (f, g, h). Then we have
𝑥1 = x + f, 𝑦1 = y + g, 𝑧1 = z + h ………………………………………………..(1)
moment of inertia of the body about OX1

5|Page
= ∑m(𝑦1 2+𝑧1 2)
= ∑m(y2+2yg+g2+z2+2hz+h2)

Figure 4

=∑m(y2+z2)+2g∑my+2h∑mz+∑m(g2+h2)…………………………………(2)
∑𝑚𝑦
But = the y coordinate of the center of inertia referred to G as origin = 0.
∑𝑚

Hence ∑m.2yg = 0, similarly ∑m.2hz = 2h.∑mz =0………………….(3)


Now from (2), we have
Moment of inertia about OX1 = ∑m(y2+z2)+∑m(g2+h2)
= ∑m(y2+z2)+(g2+h2) ∑m
= ∑m(y2+z2)+M(g2+h2)
= moment of inertia about GX + moment of inertia of a mass M placed at G about the axis OX1,
Again the product of inertia about the axis OX1 and OY1
=∑m𝑥1 𝑦1 = ∑m (x + f)(y + g)
= ∑ mxy + g ∑mx + f ∑my + fg ∑m
= ∑mxy + Mfg [by (3)]
= moment of inertia about GX and GY + product of inertia of a mass M placed at G about the
axis OX1 and OY1

Problem:
If the shape shown in Figure 5 is made by welding three 10 mm thick metal discs (each with
mass 50 kg) to a metal ring with mass 100 kg. If rotated about a central axis (out of the page),
what is the rotational inertia of the object?

6|Page
Solution:

We begin by finding the rotational inertia of the four components separately. Using the
previously given equation for a hollow cylinder, we can find the rotational inertia Ib of the big
disc. This centroid of this component already coincides with the rotational axis of the final part
so no correction is necessary.
𝟏
Ib = 𝟐m(ri2+ro2)
𝟏
= 𝟐(100kg).(0.752+12) m2

≃ 78.125 kg.m2

Using the same procedure for the smaller discs, for now centered on their own rotational axes:
𝟏
Is = 𝟐mrc2
𝟏
= .(50kg).(0.30m)2
𝟐

= 2.25 kg.m2
Because each of these three discs are rotating about a point distance d away from their respective
centroids we can use the parallel axis theorem to find the effective rotational inertia I’s
1
d = 2(1+0.75)=0.875m

I’s = Is + md2
= (2.25 kg⋅m2)+(50 kg)⋅(0.875 m)2
≃ 40.5 kg⋅m2
As each of the three small discs are at the same radius from the rotational axis of the part, we can
treat them as one part with three times the rotational inertia. Finally, the rotational inertial of the
object is:
I ≃ 78+3⋅40.5
≃ 200 kg⋅m2

7|Page
Figure 5: A system of one large hollow disc and three smaller filled discs

Where else does moment of inertia come up in physics?

Moment of inertia or rotational inertia is important in almost all physics problems that involve
mass in rotational motion. It is used to calculate angular momentum and allows us to explain (via
conservation of angular momentum) how rotational motion changes when the distribution of
mass changes. It also is needed to find the energy which is stored as rotational kinetic energy in a
spinning flywheel.

1.10 Perpendicular axis theorem:


Where does it applicable-
This is applicable only to plane lamina. The moment of inertia of a
plane lamina about an axis perpendicular to its plane is equal to the sum of the moments of
inertia of the lamina about any two mutually perpendicular axes, passing in its own plane,
intersecting each other at the point through which the perpendicular axis passes.
The theorem states that “the moment of inertia of a plane lamina about an axis perpendicular to
its plane is equal to the sum of the moments of inertia of the lamina about any two mutually

8|Page
perpendicular axes in its plane and intersecting each other at the point where the perpendicular
axis passes through it.”
Explanation:
Let OZ be the axis perpendicular to the plane lamina and passing through point O.
Let OX and OY be two mutually perpendicular axes in the plane of the lamina and intersecting at
point O. If Ix, Iy and Iz are the moments of inertia of the plane lamina about the axes OX, OY and
OZ respectively, then accordingly to the theorem of perpendicular axes, Iz = Ix+Iy .

Proof:
Let, the lamina consists of n number of particles of masses 𝑚1 , 𝑚2 , 𝑚3 , …, mn. And their
positions are at the distances 𝑟1, 𝑟2 , 𝑟3 , …, rn respectively from the origin O. For simplicity,
Let a particle of mass 𝑚1 , is at point P (𝑥1 , 𝑦1 ) and its position is denoted by 𝑟1 .
So, 𝑟12 = 𝑥1 2 + 𝑦1 2
The moment of inertia of the particle of mass m, about OZ axis is,
I1=𝑚1 𝑟12 =𝑚1 (𝑥1 2 + 𝑦1 2)
Similarly the moment of inertia of the particle of mass m2 about OZ axes is,
I2=𝑚2 (𝑥2 2 + 𝑦2 2)
Therefore, the moment of inertia of the whole lamina about OZ axis is,
Iz = I1 + I2 + … + In
= 𝑚1 (𝑥1 2 + 𝑦1 2) + 𝑚2 (𝑥2 2 + 𝑦2 2) + … + mn(xn2 + yn2)
= (𝑚1 𝑥1 2 + 𝑚2 𝑥2 2 + … + mnxn2) + (𝑚1 𝑦1 2 + 𝑚2 𝑦2 2 + … + mnyn2)
So, Iz = Ix + Iy [Proved]

9|Page
1.11 Problems:

1. Moment of inertia of a uniform rod of length 2a and mass M about a line perpendicular
to the rod.

Let AB = 2a be the rod, AC, perpendicular to AB at A. mass per unit length of the rod is
𝑴
equal to 𝟐𝒂
.

Let P be a point at a distance x from A to Q, a point adjacent to P where AP = x, PQ = δx. The mass of
𝑴
the element δx = (𝟐𝒂) δx.

𝑴
Moment of inertia about AC of the element = (𝟐𝒂) δx.x2. C
𝒂𝑴 𝟒
Moment of inertia of the whole rod about AC = ∫𝟎 𝒙𝟐 𝒅𝒙 = 𝟑 𝑴𝒂𝟐
𝟐𝒂

Moment of inertia of a uniform rod of length 2a and mass M about a


line GC perpendicular to the rod at its middle point.
Let G be the middle point of the rod AB = 2a. Now let GP = x and
A P Q B
PQ = δx.
𝑴
Figure 6
Mass of δx = 𝟐𝒂δx

A G P Q B

Figure 7
𝑴
Limits of x are from -a to +a. Moment of inertia of this element about GC = δx.x2.
𝟐𝒂
𝒂 𝑴 𝟏
Moment of inertia of the rod AB about GC =∫−𝒂 𝟐𝒂 𝒙𝟐 𝒅𝒙 = 𝟑 𝑴𝒂𝟐

10 | P a g e
Moment of inertia of a uniform rectangular lamina of mass M about an axis through its center
parallel to one of its side.

Let ABCD be the lamina such that AB = 2a, AD = 2b, center of the
lamina is at O.
Consider an elementary strip parallel to AD at a distance x from O.
Mass of the strip
𝑀
δx = 4𝑎𝑏 δx
1 𝑀
moment of inertia of the strip about OX parallel to AB = 3 𝑏 2 2𝑎 δx Figure 8: uniform rectangular lamina

𝑀𝑏 2 𝑎
moment of inertia of the rectangle about OX = 6𝑎
∫−𝑎 𝑑𝑥
𝑀𝑏 2 1
= 2𝑎 = 3 𝑀𝑏 2
6𝑎
1
moment of inertia of the rectangle about OY parallel to AD (2b) = 3 𝑀𝑎2

2. Moment of inertia of a uniform circular ring about a diameter.


Solution:
Let O be the center of the circle and XOX’ be its diameter. Let ∠POX =  and ∠QOX =  + 
∴ arc PQ = a 
Where a is the radius of the circle. Mass of the
𝑀
arc PQ = 2𝜋𝑎a 

Where M is the mass of the circle.


𝑀 𝑀 Figure 8: uniform circular ring
Moment of inertia of it about OX = 2𝜋  PN2 = 2𝜋a2sin2 

𝑀 2𝜋 2𝑀𝑎2 1 1 1
Moment of inertia if the whole circle about OX = 2𝜋 𝑎2 ∫0 sin 2  d = 𝜋= 2 𝑀𝑎2
𝜋 22

11 | P a g e
3. Moment of inertia of a uniform circular disk about diameter.
Solution:
Let O be the center and a bet the radius of the disc. Consider two
concentric circles of radii r and r+ r.

The area between these two circles = 2 𝜋rr. Mass of this area =
𝑀
2 𝜋rr
𝜋𝑎2

𝑀 2 𝜋𝑟𝑟𝑟 2
Moment of inertia about OX = 𝜋𝑎2 2

Moment of inertia whole disc about


2𝜋𝑀 𝑎
OX = 𝜋𝑎2
∫0 𝑟 3 𝑑𝑟 = Ma2 Figure 9: uniform circular disk

4. Find the product of inertia of an elliptic quadrant about its axis.


Solution:
Let us consider an elementary area xy at P(x, y).
Let the equation of the ellipse be
𝑥2 𝑦2
+ 𝑏2 = 1
𝑎2

Now product of inertia of the elliptic quadrant about OX and OY


𝑎 𝑦 𝑎 1
∫0 ∫0 𝑥𝑦𝜌𝑑𝑥𝑑𝑦 = ∫0 𝜌𝑥 2 𝑦 2 𝑑𝑥
Figure 10: elliptic quadrant
1 𝑎 𝑥2
= ρ∫ 𝑥 𝑏 2 (1 − ) 𝑑𝑥
2 0 𝑎2

1 𝑏 2 𝑎4 𝑎4
= 2ρ𝑎2 ( 2 − )
3
1
= 3 𝜌𝑎2 𝑏 2
1
=2𝜋 𝑀𝑎
1
∴ M = 4 𝜋𝑎𝑏

12 | P a g e
CHAPTER TWO: CENTER OF GRAVITY

2.1 Introduction:
Every particle of motion is attracted to the center of the earth and the force with which the earth
attracts any particle to itself is, as we shall see in Dynamics, proportional to the mass of the
particle.
Any body may be considered as an agglomeration of particles. If the body be small, compared
with the earth, line joining its component particles to the center of the earth will be very
approximately parallel and we shall consider them to be absolutely parallel.
On every particle, therefore of a rigid body there is acting a force vertically downwards which
we call its weight.
These forces may be the process of compounding parallel forces, be compounded into a single
force, equal to the sum of the weights of the particles acting at some definite point of the body,
such a point is called the center of gravity of the body.

2.2 History:
The concept of "center of mass" in the form of the center of gravity was first introduced by the
great ancient Greek physicist, mathematician, and engineer Archimedes of Syracuse. He worked
with simplified assumptions about gravity that amount to a uniform field, thus arriving at the
mathematical properties of what we now call the center of mass. Archimedes showed that
the torque exerted on a lever by weights resting at various points along the lever is the same as
what it would be if all of the weights were moved to a single point—their center of mass. In
work on floating bodies he demonstrated that the orientation of a floating object is the one that
makes its center of mass as low as possible. He developed mathematical techniques for finding
the centers of mass of objects of uniform density of various well-defined shapes.
Later mathematicians who developed the theory of the center of mass include Pappus of
Alexandria, Guido Ubaldi, Francesco Maurolico, Federico Commandino, Simon Stevin, Luca
Valerio, Jean-Charles de la Faille, Paul Guldin, John Wallis, Louis Carré, Pierre Varignon,
and Alexis Clairaut.
Newton's second law is reformulated with respect to the center of mass in Euler's first law.

13 | P a g e
2.3 Definition:
The center of gravity of a body or system of particles rigidly connected together is that point
through which the line of action of the weight of the body always passes.

2.4 Definitive intuition:


The center of gravity of any rigid body is a point such that, if the body be conceived to be
suspended from that point being released from rest and force to rotate in all directions around
this point, the body so suspended will remain at rest and preserve its original position, no matter
what the initial orientation of the relative to the ground.

2.5 Equilibrium Toys:


A toy that is known to everyone is the roly-poly doll. It is made by using the concept of center of
gravity. To build a roly-poly we need only two hemispheres or Styrofoam spherical shells, plus
some shot modeling clay or another weight. The center of gravity of the homogeneous spheres is
located at the center of the spheres. The center of gravity of the extra weight is located at the
center of the extra weight, assuming it is spherical in shape.
When we plays the shot at the bottom of one of the hemispheres, the CG of the whole system is
located between the shot and the center of the sphere as in figure 11(a).

Figure 11(a): The roly-poly doll

This is the position of stable equilibrium for the roly-poly doll, as the CG for the whole system is
at its lowest position. By tipping the roly-poly clockwise or anticlockwise, we shift its CG away
from the vertical, passing through the new point of support, raising the CG. Gravity returns the
doll to its stable position, as in Figure 11(b).

14 | P a g e
Figure 11(b): Stable equilibrium of the roly-poly doll

The main Roly-poly doll looks like,

Figure 12: Roly-poly doll

2.6 General formulae for the determination of the center of gravity


If a system of particles whose weights are w1, W2, …, wn be at points whose coordinates referred
to fixed axis ox, oy, oz in space are (x1, y1, z1), (x2, y2, z2), …, (xn, yn, zn), then the coordinates
(𝑥̅ , 𝑦̅, ̅)
𝑧 of the center of the gravity G are given by
𝑤1 𝑥1 +𝑤2 𝑥2 +⋯+𝑤𝑛 𝑥𝑛 ∑(𝑤1 𝑥1 )
𝑥̅ = =
𝑤1 +𝑤2 +⋯+𝑤𝑛 ∑(𝑤1 )

𝑤1 𝑦1 +𝑤2 𝑦2 +⋯+𝑤𝑛 𝑦𝑛 ∑(𝑤1 𝑦1 )


𝑦̅ = =
𝑤1 +𝑤2 +⋯+𝑤𝑛 ∑(𝑤1 )

𝑤1 𝑧1 +𝑤2 𝑧2 +⋯+𝑤𝑛 𝑧𝑛 ∑(𝑤1 𝑧1 )


𝑧̅ = =
𝑤1 +𝑤2 +⋯+𝑤𝑛 ∑(𝑤1 )

If all the particles on a straight line, the first of these formulae gives G: if they lie in a plane, only
the first two of these formulae give G.

15 | P a g e
2.7 General formulae for the determination of the center of gravity by
integration
In the case of continuous distribution of matter, the summations can be replaced by the definite
integrals. Thus the center of gravity of the body is given by

∫ 𝑥𝑑𝑚
𝑥̅ =
∫ 𝑑𝑚

∫ 𝑦𝑑𝑚
𝑦̅ =
∫ 𝑑𝑚
Where (x,y) stands for the coordinates of center of gravity of an elementary mass dm of the given
body. The limits of the integration will be so chosen as to cover the whole body and (𝑥̅ , 𝑦̅) stands
for the coordinates of center of gravity of the whole body.

Remark: If a body is symmetrical about a line, then the center of gravity of the body always lies
on the line of the symmetry.

2.8 Formulae for the center of gravity of an arc


Let AB be an arc of the curve y = f(x) extending from x = a to x = b. Take P(x, y) and Q (x+δx, y
+ δy) any two neighboring points on the arc. Let A be a fixed point on the curve, arc AP = s and
arc PQ = δs.
Let ρ be the density of the arc of the curve at the point P:
then ρ may be considered constant from P to Q as arc PQ = δs is
very small. Hence the mass δm of the elementary arc PQ = ρδs.
Also the C. G of this elementary arc can be approximately taken
as the point P(x, y) because the point Q is very close to P and
ultimately we have to proceed to the limits as Q → P.
Thus the coordinates (𝑥̅ , 𝑦̅) of the C. G of the arc AB are down by
the formulae

∫ 𝑥𝑑𝑚 ∫ 𝑥𝜌𝑑𝑠 ∫ 𝜌𝑥𝑑𝑠 Figure 13


𝑥̅ = = =
∫ 𝑑𝑚 ∫ 𝜌𝑑𝑠 ∫ 𝜌𝑑𝑠

∫ 𝑦𝑑𝑚 ∫ 𝑦𝜌𝑑𝑠 ∫ 𝜌𝑦𝑑𝑠


𝑦̅ = = =
∫ 𝑑𝑚 ∫ 𝜌𝑑𝑠 ∫ 𝜌𝑑𝑠

The limit of the above integrals extending from one end to the other of the arc considered.

16 | P a g e
If the equation of the curve is in cartesian coordinates
i.e. y = f(x) or x = f(y), then

𝑑𝑦
ds = √1 + ( )2 dx
𝑑𝑥

𝑑𝑥
ds = √1 + ( )2 dy
𝑑𝑦

if the equation of the curve is in polar coordinates i.e. r = f(θ), then

∫ 𝜌𝑟𝑐𝑜𝑠𝜃𝑑𝑠
𝑥̅ =
∫ 𝜌𝑑𝑠

∫ 𝜌𝑟𝑠𝑖𝑛𝜃𝑑𝑠
𝑦̅ =
∫ 𝜌𝑑𝑠

𝑑𝑟 𝑑𝜃
Where ds = √𝑟 2 + ( )2 dθ, ds = √𝑟 2 + ( )2 dr
𝑑𝜃 𝑑𝑟

If the equation of the curve are in the parametric form

𝑑𝑥 𝑑𝑦
X= f(t), y = g(t), then ds = √( )2 + ( )2 dt
𝑑𝑡 𝑑𝑡

In the case of uniform distribution (or density), ρ is constant

∫ 𝑥𝑑𝑠
𝑥̅ =
∫ 𝑑𝑠

∫ 𝑦𝑑𝑠
𝑦̅ =
∫ 𝑑𝑠

17 | P a g e
2.9 Why does center of gravity important?
The answer to this is simply that it vastly simplifies many calculations involving gravitation and
dynamics to be able to treat the mass of an object as if it is concentrated at one point called the
center of gravity or (almost, but not quite the same), the center of mass. However, it should be
made clear that the center of mass isn’t really a physical phenomenon in its own right, but more a
mathematical simplification of the physical phenomena. For example, it’s very easy to show that,
for a spherical body which is radially symmetrical in mass distribution, that gravity appears to
operate as if the mass is all concentrated at the center. However, the reality is that every point
within the sphere is attracting an external (and, for that matter, internal) mass separately.

Even irregular bodies can be shown as if they act with a mass concentrated at a single point, at
least from a distance as a good approximation; once you get very close to an irregular body, there
is no fixed center of gravity.

So the real importance of these “center” is that they vastly simplify calculations. Indeed, without
such simplifications many physical and engineering calculations would become impossibly
complex.

2.10 Center of gravity of different dimensional shapes:

One-Dimensional Figures:

Center of gravity of a straight line –

The center of gravity of any straight line is the point of bisection of the straight line.

A B

CG
Figure 14: Center of gravity of straight line

18 | P a g e
Two-Dimensional Figures:
Center of gravity of a parallelogram-
The center of gravity of a parallelogram is the point of intersection of its diagonals.

CG
Figure 15: center of gravity of a parallelogram

Center of gravity of any triangle-


The center of gravity of any triangle is at the intersection of the lines drawn from any two angles
to the middle points of the opposite sides respectively.

CG

Figure 16: center of gravity of triangle

19 | P a g e
Three dimensional figures:
Center of gravity of any cylinder-

The center of gravity of any cylinder is the point of bisection of the axis.

CG
Figure 17: center of gravity of cylinder

2.11 The Difference Between Centre of Mass and Centre of Gravity:


Many people assume that the terms “center of mass” and “center of gravity” are synonymous,
but this is not the case.

Centre of mass is the point at which the distribution of mass is equal in all directions, and does
not depend on gravitational field. Centre of gravity is the point at which the distribution of
weight is equal in all directions, and does depend on gravitational field.

Figure 18: A toy bird balances when a pivot is placed at its center of gravity.

20 | P a g e
The center of mass and the center of gravity of an object are in the same position if the
gravitational field in which the object exists is uniform. In most cases this is true to a very good
approximation: even at the top of Mount Everest (8848 meters) the gravitational field strength is
still 99.6% of its standard value. You are unlikely ever to experience a difference between center
of mass and center of gravity, as the gravitational field in which you find yourself is extremely
uniform.

But if the gravitational field strength were greater towards your feet and weaker towards your
head, then your center of gravity would be below your center of mass, perhaps somewhere
around your knees. If the gravitational field strength were greater towards your head, and weaker
towards your feet, then your center of gravity would be above your center of mass, perhaps
somewhere around your shoulders.

Figure 19:centre of gravity and center of mass of two objects

The object on the left, in a uniform gravitational field, has overlapping centers of gravity and
mass. For the object on the right, in which the gravitational field is stronger towards its base, the
center of gravity is below the center of mass. Approaching a black hole the gradient of the
gravitational field would be infinitely “steep”, leading to an incredible difference in gravitational
field and death by spaghettification for anyone falling into a black hole.

21 | P a g e
2.12 Problems:
1. Find the position of the center of gravity of the arc of the astroid
𝟐 𝟐 𝟐
𝒙𝟑 + 𝒚𝟑 = 𝒂𝟑 in the first quadrant.

Solution:
Y
Let (𝑥̅ , 𝑦̅) be the coordinates of C.G of the arc AB.
B
∫ 𝑥𝜌𝑑𝑠
𝑥̅ =
∫ 𝜌𝑑𝑠

∫ 𝑦𝜌𝑑𝑠
𝑦̅ =
∫ 𝜌𝑑𝑠
C A X

D
Limits of x and y are from 0 to a.
Figure 20: astroid
2 1
𝑑𝑠 𝑑𝑦 𝑦3 𝑎3
= √1 + ( )2 = √1 + ( )2 = 2 1
𝑑𝑥 𝑑𝑥
𝑥3 𝑥3

1 1
𝑎 −
𝑎3 ∫0 𝑥.𝑥 3 𝑑𝑥 3 2 2
𝑥̅ = 1 1 = . 𝑎= 𝑎
𝑎 − 5 3 5
𝑎3 ∫0 𝑥 3 𝑑𝑥

1 1
𝑎 −
𝑎3 ∫0 𝑦.𝑦 3 𝑑𝑦 2
𝑦̅ = 1 1 = 𝑎
𝑎 − 5
𝑎3 ∫0 𝑦 3 𝑑𝑦
2
∴ 𝑥̅ = 𝑦̅ = 𝑎
5

22 | P a g e
2. Find the position of the C.G of the arc of a semi-cardioid,
r = a (1 + cos θ).
Solution:
𝜃
Let r = a (1 + cos θ) = 2a 𝑐𝑜𝑠 2 2

Let (𝑥̅ , 𝑦̅) be the coordinates of C.G of the area .


Limit of θ is 0 to π
𝑑𝑠
∫ 𝑥𝜌𝑑𝑠 ∫ 𝑟𝑐𝑜𝑠𝜃 𝑑𝜃𝑑𝜃
𝑥̅ = = 𝑑𝑠
∫ 𝜌𝑑𝑠 ∫𝑑𝜃𝑑𝜃

𝑑𝑠 𝑑𝑟
But = √𝑟 2 + ( )2
𝑑𝜃 𝑑𝜃

=a√{(1 + 𝑐𝑜𝑠𝜃)2 + 𝑠𝑖𝑛2 𝜃}


1 Figure 21: semi-cardioid
= 2𝑎𝑐𝑜𝑠 2 𝜃
2
𝜋 1 1 𝜋 1
4𝑎 ∫0 𝑎𝑐𝑜𝑠2 2𝜃 cos 𝜃𝑐𝑜𝑠2 2𝜃𝑑𝜃 2𝑎2 ∫0 cos 𝜃𝑐𝑜𝑠2 2𝜃𝑑𝜃 16 𝑎2 4
𝑥̅ = 𝜋 1 = = = 𝑎
2𝑎2 ∫0 𝑐𝑜𝑠2 𝜃𝑑𝜃 2𝑎.2 5 4𝑎 5
2
𝜋 𝑑𝑠 𝜋 1 1 𝜋 1 1 1
∫0 𝑟𝑠𝑖𝑛𝜃 𝑑𝜃𝑑𝜃 4𝑎2 ∫0 𝑠𝑖𝑛𝜃𝑐𝑜𝑠2 2𝜃𝑐𝑜𝑠2𝜃𝑑𝜃 2𝑎 ∫0 𝑐𝑜𝑠3 2𝜃2𝑠𝑖𝑛2𝜃𝑐𝑜𝑠2𝜃𝑑𝜃
𝑦̅ = 𝜋 𝑑𝑠 = . 𝜋 1 =
∫0 𝑑𝜃 2𝑎 ∫0 𝑐𝑜𝑠2 2𝜃𝑑𝜃 2
𝑑𝜃
𝜋 1 𝜃 4𝑎
= 4𝑎 ∫0 cos4 𝜃 sin 𝑑𝜃 =
2 2 5

23 | P a g e
3. Determine the coordinate of the center of gravity of the object as shown in the figure
below.

Solution:

Figure 22
Divide the object into three parts.

Area of part 1 (A1) = (2)(6) = 12 cm2


The center point lies on the x axis (x1) = 1/2 (2) = 1 cm
The center point lies on the y axis (y1) = 1/2 (6) = 3 cm

Area of part 2 (A2) = (4)(2) = 8 cm2


The center point lies on the x axis (x2) = 2 + (1/2)(4) = 2 + 2 = 4 cm
The center point lies on the y axis (y2) = 2 + (1/2)(2) = 2 + 1 = 3 cm

Area of part 3 (A3) = (2)(6) = 12


The center point lies on the x axis (x3) = 2 + 4 + (1/2)(2) = 2 + 4 + 1 = 7 cm
The center point lies on the y axis (y3) = 1/2 (6) = 3 cm

Coordinate of the center of gravity at x axis:

24 | P a g e
Coordinate of the center of gravity at y axis:

Coordinate of the center of gravity of the object is at x axis and y axis (x, y) = (4,3)

2. Determine the coordinate of the center of gravity of object, about the x axis.

Solutions:

Figure 23

Divide the object into three parts, A, B, C, and D.


Calculate the area of each part :

25 | P a g e
AA = ½ (base)(height) = ½ (1.5)(3) = (1.5)(1.5) = 2.25
AB = (length)(width) = (4.5-1.5)(1) = (3)(1) = 3
AC = ½ (base)(height) = ½ (6-4.5)(3) = (1.5)(1.5) = 2.25
AD = ½ (base)(height) = ½ (4.5-1.5)(6-3) = ½ (3)(3) = (1.5)(3) = 4.5

Figure 24

𝑦𝐴 = 1/3 (3) = 1
𝑦𝐵 = 1/2 (1) = 0.5
𝑦𝐶 = 1/3 (3) = 1
𝑦𝐷 = 3 + (1/3) (6-3) = 3 + (1/3)(3) = 3 + 1 = 4
Coordinate of the center of gravity at y axis :

Coordinate of the center of gravity about the x axis is 2 cm.

26 | P a g e
CHAPTER THREE: VIRTUAL WORK

Virtual work arises in the application of the principle of least action to the study of forces and
movement of a mechanical system. The work of a force acting on a particle as it moves along a
displacement will be different for different displacements. Among all the possible displacements
that a particle may follow, called virtual displacements, one will minimize the action. This
displacement is therefore the displacement followed by the particle according to the principle of
least action. The work of a force on a particle along a virtual displacement is known as the
virtual work.

3.1 History
The principle of virtual work had always been used in some form since antiquity in the study of
statics. It was used by the Greeks, medieval Arabs and Latins, and Renaissance Italians as "the
law of lever". The idea of virtual work was invoked by many notable physicists of the 17th
century, such as Galileo, Descartes, Torricelli, Wallis, and Huygens, in varying degrees of
generality, when solving problems in statics. Working with Leibnizian concepts, Johann
Bernoulli systematized the virtual work principle and made explicit the concept of infinitesimal
displacement. He was able to solve problems for both rigid bodies as well as fluids. Bernoulli's
version of virtual work law appeared in his letter to Pierre Varignon in 1715, which was later
published in Varignon's second volume of Nouvelle mécanique ou Statique in 1725. This
formulation of the principle is today known as the principle of virtual velocities and is commonly
considered as the prototype of the contemporary virtual work principles. In 1743 D'Alembert
published his Traité de Dynamique where he applied the principle of θθvirtual work, based on
Bernoulli's work, to solve various problems in dynamics. His idea was to convert a dynamical
problem into static problem by introducing inertial force. In 1768, Lagrange presented the virtual
work principle in a more efficient form by introducing generalized coordinates and presented it
as an alternative principle of mechanics by which all problems of equilibrium could be solved. A
systematic exposition of Lagrange's program of applying this approach to all of mechanics, both
static and dynamic, essentially D'Alembert’s principle, was given in his Mécanique
Analytique of 1788. Although Lagrange had presented his version of least action principle prior
to this work, he recognized the virtual work principle to be more fundamental mainly because it
could be assumed alone as the foundation for all mechanics, unlike the modern understanding
that least action does not account for non-conservative forces.

3.2 Work of a Force. A force does work when it undergoes a displacement in the direction
of its line of action. Consider, for example, the force F in Fig. 25(a) that undergoes a differential
displacement dr. If θ is the angle between the force and the displacement, then the component of
F in the direction of the displacement is F cosθ. And so the work produced by F is

dU = F dr cosθ

27 | P a g e
Figure 25: work of a force
gure 25
Notice that this expression is also the product of the force F and the component of displacement
in the direction of the force, dr cosθ, Fig. 25(b). If we use the definition of the dot product the
work can also be written as
dU = F . dr

As the above equations indicate, work is a scalar, and like other scalar quantities, it has a
magnitude that can either be positive or negative.
In the SI system, the unit of work is a joule(J), which is the work produced by a 1-N force that
displaces through a distance of 1m in the direction of the force (1 J = 1 N·m). The unit of work
in the FPS system is the foot-pound (ft. lb), which is the work produced by a 1-Ib force that
displace through a distance of 1 ft in the direction of the force.
The moment of a force has this same combination of Units; however, the concepts of moment
and work are in no way related. A moment is a vector quantity, whereas work is a scalar.

3.3 Work of a Couple Moment. The rotation of a couple moment also produces work.
Consider the rigid body in Fig.26, which is acted upon by the couple forces F and - F that
produce a couple moment M having a magnitude M = Fr.

28 | P a g e
Figure 26: work of a couple moment

gure 25
When the body undergoes the differential displacement shown, points A and B move drA and
drB to their final positions A' and B', respectively. Since drB = drA + dr', this movement can be
thought of as a translation drA, where A and B move to A' and B", and a rotation about A',
where the body rotates through the angle dθ about A. The couple forces do no work during the
translation drA, because each force undergoes the Same amount of displacement in opposite
directions, thus cancelling out the work. During rotation, however, F is displaced dr" = r dθ,
and so it does work
dU = F . dr" = F . r dθ. Since M = Fr, the work of the couple moment M is therefore
dU = M dθ
If M and dθ have the same sense, the work is positive; however, if they have the opposite sense,
the work will be negative.

3.4 Definition:
The definitions of the work of a force and a couple have been presented in terms of
actual movements expressed by differential displacements having magnitudes of dr and dθ.
Consider now an imaginary or virtual movement of a body in the static equilibrium, which
indicates a displacement or rotation that is assumed and does not actually exist. These
movements are first-order differential quantities and will be denoted by the symbols δr and δθ
(delta r and delta θ), respectively. The virtual work done by a force having a virtual displacement
δr is

δU = F cosθ δr
Similarly, when a couple undergoes a virtual rotation δθ in the plane of the couple forces, the
virtual work is
δU = M δθ
29 | P a g e
3.5 Principle of Virtual Work
The principle of virtual work states that if a body is in equilibrium, then the algebraic sum of the
virtual work done by all the forces and couple moments acting on the body, is zero for any
virtual displacement of the body. Thus,

δU = 0
For example, consider the free-body diagram of the particle (ball) that rests on the floor, fig.27.
If we imagine the ball to be displaced downwards a virtual amount δy, then the weight does
positive virtual work, W δy, and the normal force does negative virtual work, - N δy. For
equilibrium the total virtual work must be zero, so that δU = W δy – N δy = (W – N) δy = 0.
Since δy ≠ 0, then N = W as required by applying ∑ Fy =0.

Figure 27: Free body diagram of a ball

In a similar manner, we can also apply the virtual-work equation δU = 0 to a rigid body subjected
to a coplanar force system. Here, separate virtual translations in the x and y directions and a
virtual rotation about an axis perpendicular to the x-y plane that passes through an arbitrary point
O, will correspond to the three equilibrium equations ∑Fx = 0, ∑Fy = 0, and ∑M0 = 0. When
writing these equations it is not necessary to include the work done by the internal forces acting
within the body since a rigid body does not deform when subjected to an external loading, and
furthermore, when the body moves through a virtual displacement, the internal forces occur in
equal but opposite collinear pairs, so that the corresponding work done by each pair of forces
will cancel.
To demonstrate an application, consider the simply supported beam in Fig.28(a). When the beam
is given a virtual rotation δθ about point B, Fig. 28(b), the only forces that do work are P and Ay.
Since δy = l δθ and δy' = (l/2) δθ, the virtual work equation for this case is δU = Ay(l δθ) –
P(l/2)δθ = (Ayl – P l/2) δθ = 0 .Since δθ ≠ 0, then Ay = P/2. Excluding δθ, notice that the terms in
parentheses actually represent application of ∑Mb = 0.

30 | P a g e
Figure 28: Simply supported beam and its rotation

As seen from the above two examples, no added advantage is gained by solving particle and
rigid-body equilibrium problems using the principle of virtual work. This is because for each
application of the virtual-work equation, the virtual displacement, common to every term, factors
out, leaving an equation that could have been obtained in a more direct manner, by simply
applying an equation or equilibrium.

3.6 Principle of Virtual Work for a System of Connected Rigid Bodies:


The method of virtual work is particularly effective for solving equilibrium problems that involve a
system of several connected rigid bodies, such as the ones shown in Fig. 29
Each of these systems is said to have only one degree of freedom since the arrangement of the links can
be completely specified using only one coordinate θ. In other words with this single coordinate and the
length of the members, we can locate the position of the forces F and P.

Figure 29: Several connected rigid bodies

31 | P a g e
In this text, we will only consider the application of the principle of virtual work to systems containing
one degree of freedom. Because they are less complicated, they will serve as a way to approach the
solution of more complex problems involving, systems with many degrees of freedom. The procedure for
solving problems involving a system of frictionless connected rigid bodies follows.

Important Points:
 A force does work when it moves through a displacement in the direction of the force. A couple
moment does work when it moves through a collinear rotation. Specifically, positive work is done
when the force or couple moment and its displacement have the same sense of direction.

 The principle of virtual work is generally used to determine the equilibrium configuration for a
system of multiply connected members.

 A virtual displacement is imaginary; i.e., it does not really happen. It is a differential


displacement that is given in the positive direction of a position coordinate.

 Forces or couple moments that do not virtually displace do no virtual work.

3.7 Procedure for Analysis:


Free-Body Diagram
 Draw the free-body diagram of the entire system of connected bodies and define the
coordinate q.
 Sketch the “deflected position” of the system on the free-body diagram when the system
undergoes a positive virtual displacement δq.
Virtual Displacements
 Indicate position coordinates s, each measured from a fixed point on the free-body
diagram. These coordinates are directed to the forces that do work.

 Each of these coordinate axes should be parallel to the line of action of the force to which
it is directed, so that the virtual work along the coordinate axes can be calculated.

 Relate each of the position coordinates s to the coordinate q then differentiate these
expressions in order to express each virtual displacement δs in terms of δq.

32 | P a g e
Virtual-Work Equation

 Write the virtual-work equation for the system assuming that, whether possible or not,
each position coordinate s undergoes a positive virtual displacement δs. If a force or
couple moment is in the same direction as the positive virtual displacement, the work is
positive. Otherwise, it is negative.

 Express the work of each force and couple moment in the equation in terms of δq.

 Factor out this common displacement from all the terms, and solve for the unknown
force, couple moment, or equilibrium position q.

3.8 D’Alembert’s principle of virtual work:


A system of particles is in equilibrium if and only if the total virtual work of the actual forces
is zero i.e. if
∑Fi(effective) . δri = 0 …………… (1)
Which is called the principle of virtual work.
But it can be related to obtain a similar principle for dynamics. This principle of virtual work
was modified into a new principle by James Bernoulli and developed by D’Alembert by
taking
𝑑𝑝𝑖
Fi = when pi is the momentum of i-th particle due to force Fi acting on i-th particle.
𝑑𝑡
𝑑𝑝𝑖
Or, Fi +(- ) = 0 ……………………….. (2)
𝑑𝑡
𝑑𝑝
The actual together with the added force when is often called the reserved effective force
𝑑𝑡
on the particles.
By the principle of virtual work we have
∑(Fi – 𝑝𝑖̇ ). δri = 0
Which is called the D’Alembert’s principle.

33 | P a g e
3.9 Problems:

1. Determine the required force P in Fig. 30, needed to maintain equilibrium of the scissors
linkage when θ = 60⁰. The spring is unstretched when θ = 30⁰. Neglect mass of the links.

Solution:
Free-Body Diagram. Only Fs and P do work when θ undergoes a positive virtual
displacement δθ, Fig. 30. For the arbitrary position θ, the spring is stretched (0.3 m) sin θ -
(0 .3 m) sin 30⁰, so that
Fs = ks = 5000 N/m [(0.3 m) sin θ - (0 .3 m) sin 30⁰]
= (1500 sin θ - 750) N

Figure 30: equilibrium of the scissors linkage

Virtual Displacements. The position coordinates, XB and XD, measured from the fixed point
A, are used to locate Fs and P. These coordinates are parallel to the line of action of their
corresponding forces. Expressing XB and XD in terms of the angle θ using trigonometry.
XB = (0.3 m) sin θ
XD = 3[(0.3 m) sin θ

34 | P a g e
= (0.9 m) sin θ
Differentiating, we obtain the virtual displacements of the points B and D.
δXB = 0.3 cos θ δθ
(1)
δXD = 0.9 cos θ δθ
(2)
Virtual-Work Equation. Force P does positive work since it acts in the positive sense of its
virtual displacement. The spring force Fs does negative work since it acts opposite to its
positive virtual displacement. Thus, the virtual-work equation becomes
δU = 0; -Fs δXB + P δXD = 0
- [1500 sin θ – 750] (0.3 cos θ δθ ) + P (0.9 cos θ δθ ) = 0
[0.9P + 225 – 450 sin θ] cos θ δθ = 0
Since cos θ δθ ≠ 0, then this equation requires
P = 500 sin θ – 250
When θ = 60⁰
P = 500 sin 60⁰ – 250 = 183 N

2. The mechanism in Fig. 31 supports the 50-lb cylinder. Determine the angle θ for
equilibrium if the spring has an unstretched length of 2 ft when θ = 0⁰. Neglect the mass
of the members.

Solution:

Free-Body Diagram. When the mechanism undergoes a positive virtual displacement δθ,
Fig. 31(b), only Fs and the 50-Ib force do work. Since the final length of the spring is 2(1 ft
cos θ), then
Fs = ks = (200 lb/ft)(2 ft – 2ft cos θ) = (400 – 400cos θ)lb

35 | P a g e
Figure 31: Mechanism to support a cylinder

Virtual Displacements. The position coordinates XD and XE are stablished from the fixed
point A to locate Fs at D and at E. The coordinate YB, also measured from A, specifies the
position of the 50-Ib force at B. The coordinates can be expressed in terms of θ using
trigonometry.

XD = (1 ft) cos θ
XE = 3[(1 ft) cos θ] = (3 ft) cos θ
YB = (2 ft) sin θ
Differentiating, we obtain the virtual displacements of points D, E, and B as
δXD = -1 sin θ δθ (1)
δXE = -3 sin θ δθ (2)
δYB = 2 cos θ δθ (3)

Virtual-Work Equation. The virtual-work equation is written as if all virtual displacements are
positive, thus
δU = 0; Fs δXE + 50 δYB – Fs δXD = 0
(400 – 400 cos θ )( - 3 sin θ δθ) + 50(2 cos θ δθ) – (400 – 400 cos θ)(- 1 sin θ δθ) = 0
δθ (800 sin θ cos θ – 800 sin θ + 100 cos θ) = 0
since δθ ≠ 0, then
800 sin θ cos θ – 800 sin θ + 100 cos θ = 0
Solving by trial and error,

36 | P a g e
θ = 34.9⁰

37 | P a g e
3. A heavy uniform rod of length 2a rests on two smooth planes inclined at angles α and β
to the horizon. If θ be the inclination of the rod to the horizon, then prove that
𝟏
tan θ = 𝟐(cot α – cot β)

Solution:
Let the rod AB rests on two smooth planes
which are inclined at angles α and β to the
horizon.
Let G be the mid-point of AB. Since given AB =
2a, so AG = GB = a. Draw AL, GN and BM
perpendiculars to the horizon through P. Also
draw AO parallel to the horizontal line LM.
Then AL = ON.
Now ∠APL = α and ∠BPM = β
Let W be the weight of the rod AB which acts at
G vertically downwards.
Figure 32: An inclined uniform rod
Then the virtual work done by W is
W. δ(GN) = 0
Since W ≠ 0, so δ(GN) = 0
But GN = GO + ON = GA sin θ + AL = a sin θ + AP sin α
From the properties of the triangle ABP, we get
𝐴𝑃 𝐴𝐵 𝑃𝐵
= =
sin 𝑃𝐵𝐴 sin 𝐴𝑃𝐵 sin 𝑃𝐴𝐵
𝐴𝑃 2𝑎 2𝑎
Or, sin(𝛽 − 𝜃) = sin{𝜋−(𝛼 + 𝛽)} = sin(𝛼 + 𝛽)
sin(𝛽 − 𝜃)
∴ AP = 2a . sin(𝛼 + 𝛽)
sin(𝛽 − 𝜃)
Therefore, GN = a sin θ + 2a sin(𝛼 + 𝛽) . sin α

−cos (𝛽 − 𝜃)
∴ δ(GN) = a cos θ δθ + 2a sin α δθ = 0
sin(𝛼 + 𝛽)
2cos (𝛽 − 𝜃)
Or, cos θ - sin α = 0
sin(𝛼 + 𝛽)

Or, sin (α + β) cos θ – 2 cos ( β – θ) sin α = 0

Or, (sin α cos β + cos α sin β) cos θ = 2(cos β cos θ + sin β sin θ) sin α

38 | P a g e
Or, (cos α sin β - sin α cos β) cos θ = 2 sin α sin β sin θ
1 (cos α sin β − sin α cos β) sin 𝜃
Or, 2 ( )=
sin α sinβ cos 𝜃
1
Or, 2 (cot α – cot β) = tan θ
1
∴ tan θ = 2 (cot α – cot β)

3. Two equal uniform rods AB and AC, each of length 2b, are freely joined at A and
rests on a smooth vertical circle of radius a. So that if 2θ be the angle between them,
then b sin3 θ = a cos θ.
Solution: Let W be the weight of each rod acting at the
mid-points G1 and G2 vertically downwards.
Then AG1 = AG2 = b, since AB = AC = 2b
Let D and E be the points of the rods AB and AC
respectively which are in contact with the circle whose
center is O. Then OD = OE = a.
Join A and O. Then ∠DAO = ∠EAO = θ
Also join G1 and G2. The line G1 G2 intersects AO in G
(say).
Then G1G = G2G
Now OG = OA - AG.... (1) Figure 33: Two joined uniform rods

From the right angled triangle AOE, we have


𝑂𝐸 𝑎
Sin θ = 𝐴𝑂 = ∴ OA = a cosec θ
𝐴𝑂

Again from the right angled triangle AG G2, we have


𝐴𝐺 𝐴𝐺
Cos θ = =
𝐴𝐺2 𝑏

∴ AG = b cos θ
putting the values of OA and AG in (1)
𝑎
we get OG = a cosec θ – b cos θ = sin θ − 𝑏 cos θ

39 | P a g e
Here 2W is the only force that comes into the equation of the virtual work and OG the height of
the center of gravity of each rod above the center of the circle.
Now by the principle of virtual work we have
2W. δ(OG) = 0
or, δ(OG) = 0. since 2W ≠ 0.
𝑎
or, δ(sin 𝜃 − 𝑏 cos 𝜃) = 0
𝑎
or,− 𝑠𝑖𝑛2 𝜃 𝑐𝑜𝑠𝜃 + 𝑏 sin 𝜃 = 0

or, - a cos θ + b sin3 θ = 0


or, b sin3 θ = a cos θ

40 | P a g e
CHAPTER FOUR: EQUATIONS OF MOTION

4.1 History
Historically, equations of motion first appeared in classical mechanics to describe the motion
of massive objects, a notable application was to celestial mechanics to predict the motion of the
planets as if they orbit like clockwork (this was how Neptune was predicted before its
discovery), and also investigate the stability of the solar system.
It is important to observe that the huge body of work involving kinematics, dynamics and the
mathematical models of the universe developed in baby steps – faltering, getting up and
correcting itself – over three millennia and included contributions of both known names and
others who have since faded from the annals of history.
In antiquity, notwithstanding the success of priests, astrologers and astronomers in predicting
solar and lunar eclipses, the solstices and the equinoxes of the Sun and the period of the Moon,
there was nothing other than a set of algorithms to help them. Despite the great strides made in
the development of geometry made by Ancient Greeks and surveys in Rome, we were to wait for
another thousand years before the first equations of motion arrive.
The exposure of Europe to the collected works by the Muslims of the Greeks, the Indians and the
Islamic scholars, such as Euclid’s Elements, the works of Archimedes, and Al-Khwārizmī's
treatises [3] began in Spain, and scholars from all over Europe went to Spain, read, copied and
translated the learning into Latin. The exposure of Europe to Arabic numerals and their ease in
computations encouraged first the scholars to learn them and then the merchants and invigorated
the spread of knowledge throughout Europe.
By the 13th century the universities of Oxford and Paris had come up, and the scholars were now
studying mathematics and philosophy with lesser worries about mundane chores of life—the
fields were not as clearly demarcated as they are in the modern times. Of these, compendia and
redactions, such as those of Johannes Campanus, of Euclid and Aristotle, confronted scholars
with ideas about infinity and the ratio theory of elements as a means of expressing relations
between various quantities involved with moving bodies. These studies led to a new body of
knowledge that is now known as physics.
Of these institutes Merton College sheltered a group of scholars devoted to natural science,
mainly physics, astronomy and mathematics, of similar in stature to the intellectuals at the
University of Paris. Thomas Bradwardine, one of those scholars, extended Aristotelian quantities
such as distance and velocity, and assigned intensity and extension to them. Bradwardine
suggested an exponential law involving force, resistance, distance, velocity and time. Nicholas
Oresme further extended Bradwardine's arguments. The Merton school proved that the quantity
of motion of a body undergoing a uniformly accelerated motion is equal to the quantity of a
uniform motion at the speed achieved halfway through the accelerated motion.
For writers on kinematics before Galileo, since small time intervals could not be measured, the
affinity between time and motion was obscure. They used time as a function of distance, and in

41 | P a g e
free fall, greater velocity as a result of greater elevation. Only Domingo de Soto, a Spanish
theologian, in his commentary on Aristotle's Physics published in 1545, after defining "uniform
difform" motion (which is uniformly accelerated motion) – the word velocity wasn't used – as
proportional to time, declared correctly that this kind of motion was identifiable with freely
falling bodies and projectiles, without his proving these propositions or suggesting a formula
relating time, velocity and distance. De Soto's comments are shockingly correct regarding the
definitions of acceleration (acceleration was a rate of change of motion (velocity) in time) and
the observation that during the violent motion of ascent acceleration would be negative.
Discourses such as these spread throughout Europe and definitely influenced Galileo and others,
1
and helped in laying the foundation of kinematics. Galileo deduced the equation s = 2gt2 in his
work geometrically, using the Merton rule, now known as a special case of one of the equations
of kinematics. He couldn't use the now-familiar mathematical reasoning. The relationships
between speed, distance, time and acceleration was not known at the time.
Galileo was the first to show that the path of a projectile is a parabola. Galileo had an
understanding of centrifugal force and gave a correct definition of momentum. This emphasis of
momentum as a fundamental quantity in dynamics is of prime importance. He measured
momentum by the product of velocity and weight; mass is a later concept, developed by Huygens
and Newton. In the swinging of a simple pendulum, Galileo says in Discourses that "every
momentum acquired in the descent along an arc is equal to that which causes the same moving
body to ascend through the same arc." His analysis on projectiles indicates that Galileo had
grasped the first law and the second law of motion. He did not generalize and make them
applicable to bodies not subject to the earth's gravitation. That step was Newton's contribution.
The term "inertia" was used by Kepler who applied it to bodies at rest. (The first law of motion is
now often called the law of inertia.)
Galileo did not fully grasp the third law of motion, the law of the equality of action and reaction,
though he corrected some errors of Aristotle. With Stevin and others Galileo also wrote on
statics. He formulated the principle of the parallelogram of forces, but he did not fully recognize
its scope.
Galileo also was interested by the laws of the pendulum, his first observations of which were as a
young man. In 1583, while he was praying in the cathedral at Pisa, his attention was arrested by
the motion of the great lamp lighted and left swinging, referencing his own pulse for time
keeping. To him the period appeared the same, even after the motion had greatly diminished,
discovering the isochronism of the pendulum.
More careful experiments carried out by him later, and described in his Discourses, revealed the
period of oscillation varies with the square root of length but is independent of the mass the
pendulum.
Thus we arrive at René Descartes, Isaac Newton, Gottfried Leibniz, et al.; and the evolved forms
of the equations of motion that begin to be recognized as the modern ones.
Later the equations of motion also appeared in electrodynamics, when describing the motion of
charged particles in electric and magnetic fields, the Lorentz force is the general equation which
serves as the definition of what is meant by an electric field and magnetic field. With the advent
of special relativity and general relativity, the theoretical modifications to spacetime meant the
classical equations of motion were also modified to account for the finite speed of light,
42 | P a g e
and curvature of spacetime. In all these cases the differential equations were in terms of a
function describing the particle's trajectory in terms of space and time coordinates, as influenced
by forces or energy transformations.
However, the equations of quantum mechanics can also be considered "equations of motion",
since they are differential equations of the wavefunction, which describes how a quantum state
behaves analogously using the space and time coordinates of the particles. There are analogs of
equations of motion in other areas of physics, for collections of physical phenomena that can be
considered waves, fluids, or fields.

In this section we propose to consider the production of motion, and it will be necessary to
commence with a few elementary definitions.

4.2 Related definitions


Matter is “that which can be perceived by the senses” or “that which can be acted upon by, or
can exert, force.” No definition can however be given that would convey an idea of what matter
is to anyone who did not already possess that idea. It, like time and space, is a primary
conception.
A Body is a portion of matter which is bounded by surfaces, and which is limited in every
direction.
A Particle is a portion of matter which is infinitely small in all its dimensions.
The Mass of a body is the quantity of matter in the body.
Force is that which changes, or tends to change, the state of rest or uniform motion of a body.

Intuition:
If we have a small portion of any substance, say iron, resting on a smooth table, we may by a
push be able to move it fairly easily; if we take a larger quantity of the same iron, the same effort
on our part will be able to move it less easily. Again, if we take two portions of platinum and
wood of exactly the same size and shape, the effect produced on these two substances by equal
efforts on our part will be quite different. Thus common experience shows us that the same effort
applied to different bodies, under seemingly the same conditions, does not always produce the
same result. This is because the masses of the bodies are different.

If to the same mass we apply two forces in succession, and they generate the same velocity in the
same time, the forces are said to be equal.

43 | P a g e
If the same force be applied to two different masses, and if it produce in them the same velocity
in the same time, the masses are said to be equal.
we here assume that it is possible to create forces of equal intensity on different occasions, e.g.
we assume that the force necessary to keep a spiral spring stretched through the same distance is
the same when other conditions are unaltered.

Density. The density of a uniform body is the mass of a unit volume of the body; so that, if m be
the mass of volume V of a body whose density is ρ, then
𝑚
ρ= 𝑣

The Weight of a body is the force with which the earth attracts the body.

The Momentum of a body is proportional to the product of the mass and the velocity of the
body. If we take as the unit of momentum the momentum of a unit mass moving with unit
velocity, then the momentum of a body is mv, where m is the mass and v the velocity of the
body. The direction of the momentum is the same as that of the velocity.

4.3 Newton’s laws of motion with explanations:


We can now enunciate what are commonly called Newton s Laws of Motion.
They are,

Law I. Everybody continues in its state of rest, or of uniform motion in a straight line, except
in so far as it be compelled by external impressed force to change that state.

Law II. The rate of change of momentum is proportional to the impressed force, and takes
place in the direction of the straight line in which the force acts.

Law III. When two bodies interact, the forces on the bodies from each other are always equal
in magnitude and opposite in direction.

44 | P a g e
Explanations:
Law I. We never see this law practically exemplified in nature because it is impossible ever to
get rid of all forces during the motion of the body. It may be seen approximately in operation in
the case of a piece of dry, hard ice projected along the surface of dry, well swept, ice. The only
forces acting on the fragment of ice, in the direction of its motion, are the friction between the
two portions of ice and the resistance of the air. The smoother the surface of the ice the further
the small portion will go, and the less the resistance of the air the further it will go. The above
law asserts that if the ice were perfectly smooth and if there were no resistance of the air and no
other forces acting on the body, then it would go on forever in a straight line with uniform
velocity.
The law states a principle sometimes called the Principle of Inertia, viz. that a body has no
innate tendency to change its state of rest or of uniform motion in a straight line. If a portion of
metal attached to a piece of string be swung round on a smooth horizontal table, then, if the
string break, the metal, having no longer any force acting on it, proceeds to move in a straight
line, viz. the tangent to the circle at the point at which its circular motion ceased.
If a man step out of a rapidly moving train he is generally thrown to the ground; his feet on
touching the ground are brought to rest; but, as no force acts on the upper part of his body, it
continues its motion as before, and the man falls to the ground. If a man be riding on a horse
which is galloping at a fairly rapid pace and the horse suddenly stops, the rider is in danger of
being thrown over the horse s head.
If a man be seated upon the back seat of a dog-cart, and the latter suddenly start, the man is very
likely to be left behind.

Law II. From this law we derive our method of measuring force. Let m be the mass of a body,
and F the acceleration produced in it by the action of a force whose measure is P.
Then, by the second law of motion,
P ∝ rate of change of momentum,
∝ rate of change of mv,
∝ m X rate of change of v (m being unaltered),
∝ m.F.
∴ P =k . mF, where k is some constant.
Now let the unit of force be so chosen that it may produce in unit mass the unit of acceleration.
Hence, when m = 1 and F = 1, we have P = 1,

45 | P a g e
and therefore k = 1.
The unit of force being thus chosen, we have
P= mF
Therefore, when proper units are chosen, the measure of the force is equal to the measure of the
rate of change of the momentum.

Law III. Two bodies are said to interact when they push or pull on each other—that is, when a
force acts on each body due to the other body. For example, suppose you position a book B so it
leans against a crate C (Fig. 34(a)). Then the book and crate interact: There is a horizontal force
FBC on the book from the crate (or due to the crate) and a horizontal force FCB on the crate from
the book (or due to the book).This pair of forces is shown in Fig.34 (b).

Figure 34: Interaction of two bodies

For the book and crate, we can write this law as the scalar relation
FBC = FCB (equal magnitudes)
or as the vector relation
FBC = - FCB (equal magnitudes and opposite directions),

46 | P a g e
where the minus sign means that these two forces are in opposite directions. We can call the
forces between two interacting bodies a third-law force pair. When any two bodies interact in
any situation, a third-law force pair is present. The book and crate in Fig. 34(a) are stationary,
but the third law would still hold if they were moving and even if they were accelerating.

As another example, let us find the third-law force pairs involving the cantaloupe in Fig. 35(a),
which lies on a table that stands on Earth. The cantaloupe interacts with the table and with Earth
(this time, there are three bodies whose interactions we must sort out).
Let’s first focus on the forces acting on the cantaloupe (Fig. 35(b)). Force FCT is the normal force
on the cantaloupe from the table, and force FCE is the gravitational force on the cantaloupe due to
Earth. Are they a third-law force pair? No, because they are forces on a single body, the
cantaloupe, and not on two interacting bodies.

Figure 35: Interaction of a cantaloupe with a table and the earth

To find a third-law pair, we must focus not on the cantaloupe but on the interaction between the
cantaloupe and one other body. In the cantaloupe– Earth interaction (Fig. 35(c)), Earth pulls on
the cantaloupe with a gravitational force FCE and the cantaloupe pulls on Earth with a
gravitational force FEC. Are these forces a third-law force pair? Yes, because they are forces on
two interacting bodies, the force on each due to the other. Thus, by Newton’s third law,
FCE = - FEC (cantaloupe-Earth interaction).
Next, in the cantaloupe–table interaction, the force on the cantaloupe from the table is FCT and,
conversely, the force on the table from the cantaloupe is FTC (Fig. 35(d)).These forces are also a
third-law force pair, and so
FCT = - FTC (cantaloupe-table interaction).

47 | P a g e
4.4 One Dimensional Motion:
4.4.1 Uniform acceleration
The differential equation of motion for a particle of constant or uniform acceleration in a straight
line is simple: the acceleration is constant, so the second derivative of the position of the object is
constant. The results of this case are summarized below.
4.4.2 Constant translational acceleration in a straight line
These equations apply to a particle moving linearly, in three dimensions in a straight line with
constant acceleration. Since the position, velocity, and acceleration are collinear (parallel, and lie
on the same line) – only the magnitudes of these vectors are necessary, and because the motion is
along a straight line, the problem effectively reduces from three dimensions to one.

where:

 r0 is the particle's initial position


 r is the particle's final position
 v0 is the particle's initial velocity
 v is the particle's final velocity
 a is the particle's acceleration
 t is the time interval

Here a is constant acceleration, or in the case of bodies moving under the influence of gravity,
the standard gravity g is used. Note that each of the equations contains four of the five variables,
so in this situation it is sufficient to know three out of the five variables to calculate the
remaining two.
In elementary physics the same formulae are frequently written in different notation as:

48 | P a g e
where u has replaced 𝑣0 ., s replaces r, and 𝑠0 = 0. They are often referred to as the SUVAT
equations, where "SUVAT" is an acronym from the variables: s = displacement (𝑠0 . = initial
displacement), u = initial velocity, v = final velocity, a = acceleration, t = time.

4.4.3 Related problems:

A block slides along a frictionless surface with a constant acceleration of 2 m/s2. At time t = 0 s
the block is at x = 5m and travelling with a velocity of 3 m/s.
a) Where is the block at t = 2 seconds?
b) What is the block’s velocity at 2 seconds?
c) Where is the block when it’s velocity is 10 m/s?
d) How long did it take to get to this point?

Solution:
Here is an illustration of the setup.

Figure 36: A block slides along a frictionless surface

49 | P a g e
The variables we know are:
𝑥0 = 5 m
𝑣0 . = 3 m/s
a = 2 m/s2

Part a) Where is the block at t = 2 seconds?


Equation 1 is the useful equation for this part.
x = 𝑥0 + 𝑣0 .t + ½at2
Substitute t = 2 seconds for t and the appropriate values of 𝑥0 and 𝑣0 .
x = 5 m + (3 m/s)(2 s) + ½(2 m/s2)(2 s)2
x=5m+6m+4m
x = 15 m
The block is at the 15 meter mark at t = 2 seconds.

Part b) What is the block’s velocity at t = 2 seconds?


This time, Equation 2 is the useful equation.
v = 𝑣0 . + at
v = (3 m/s) + (2 m/s2)(2 s)
v = 3 m/s + 4 m/s
v = 7 m/s
The block is travelling 7 m/s at t = 2 seconds.

Part c) Where is the block when it’s velocity is 10 m/s?


Equation 3 is the most useful at this time.
v2 = 𝑣0 .2 + 2a(x – 𝑥0 .)
(10 m/s)2 = (3 m/s)2 + 2(2 m/s2)(x – 5 m)
100 m2/s2 = 9 m2/s2 + 4 m/s2(x – 5 m)
91 m2/s2 = 4 m/s2(x – 5 m)
22.75 m = x – 5 m
27.75 m = x
The block is at the 27.75 m mark.

Part d) How long did it take to get to this point?


There are two ways you could do this. You could use Equation 1 and solve for t using the value

50 | P a g e
you calculated in part c of the problem, or you could use equation 2 and solve for t. Equation 2 is
easier.
v = 𝑣0 . + at
10 m/s = 3 m/s + (2 m/s2)t
7 m/s = (2 m/s2)t
7⁄2 s = t
It takes 7⁄2 s or 3.5 s to get to the 27.75 m mark.

4.5 Two Dimensional Motion

Figure 37: A spacecraft equipped with two engines

To understand how displacement, velocity, and acceleration are applied to two-dimensional


motion, consider a spacecraft equipped with two engines that are mounted perpendicular to each
other. These engines produce the only forces that the craft experiences, and the spacecraft is
assumed to be at the coordinate origin when t0 = 0 s, so that r0 = 0 m. At a later time t, the
spacecraft’s displacement is Dr = r–r0 = r. Relative to the x and y axes, the displacement r has
vector components of x and y, respectively.
In Figure only the engine oriented along the x direction is firing, and the vehicle accelerates
along this direction. It is assumed that the velocity in the y direction is zero, and it remains zero,
since the y engine is turned off. The motion of the spacecraft along the x direction is described by
the five kinematic variables x, ax, vx, v0x, and t. Here the symbol “x” reminds us that we are
dealing with the x components of the displacement, velocity, and acceleration vectors. The
variables x, ax, vx, and v0x are scalar components. These components are positive or negative

51 | P a g e
numbers (with units), depending on whether the associated vector components point along the
+x or the –x axis.

Equations of Kinematics for Constant Acceleration Two-Dimensional Motion

4.5.1 Projectile motion:

We next consider a special case of two-dimensional motion: A particle moves in a vertical plane
with some initial velocity but its acceleration is always the freefall acceleration, which is
downward. Such a particle is called a projectile (meaning that it is projected or launched), and its
motion is called projectile motion.

There are a variety of examples of projectiles. An object dropped from rest is a projectile
(provided that the influence of air resistance is negligible). An object that is thrown vertically
upward is also a projectile (provided that the influence of air resistance is negligible). And an
object which is thrown upward at an angle to the horizontal is also a projectile (provided that the
influence of air resistance is negligible). A projectile is any object that once projected or dropped
continues in motion by its own inertia and is influenced only by the downward force of gravity.

52 | P a g e
Figure 38: Projectile motion

4.5.2 Velocity, Acceleration and Displacement of the Projectile:

Consider a projectile being launched at an initial velocity v0 in a direction making an angle θ


with the horizontal. We assume that air resistance is negligible and the only force acting on the
object is the force of gravity with acceleration g = 9.8 m/s2.

Figure 39: Projectile motion and its graph

The initial velocity V0 being a vector quantity, has two components:


V0x and V0y given by

53 | P a g e
V0x = V0 cos(θ)
V0y = V0 sin(θ)
The acceleration A is a also a vector with two components Ax and Ay given by
Ax = 0 and Ay = - g = - 9.8 m/s2
Along the x axis the acceleration is equal to 0 and therefore the velocity Vx is constant and is
given by
Vx = V0 cos(θ)
Along the y axis, the acceleration is uniform and equal to - g and the velocity at time t is given
by
Vy = V0 sin(θ) - g t
Along the x axis the velocity Vx is constant and therefore the component x of the displacement is
given by
x = V0 cos(θ) t
Along the y axis, the motion is that of a uniform acceleration type and the y component of the
displacement is given by
y = V0 sin(θ) t - (1/2) g t2

We can find the equation of the projectile’s path (its trajectory) by eliminating time t. Then we
obtain, after a little rearrangement,
𝑔𝑥 2
Y = (tan θ) x - 2(𝑣 (trajectory)
0 𝑐𝑜𝑠(𝜃))2

This is of the form y = ax + bx2, in which a and b are constants. This is the equation of a
parabola, so the path is parabolic.

Horizontal range:
The horizontal range R of the projectile is the horizontal distance the projectile has traveled
when it returns to its initial height (the height at which it is launched).To find range R, let us put
x – x0 = R and y – y0 = 0, we obtain
2𝑣02
R= sin 𝜃 cos 𝜃
𝑔

𝑣02
= sin 2𝜃
𝑔

54 | P a g e
4.5.3 Problems:
Figure shows a pirate ship 560 m from a fort defending a harbor entrance. A defense cannon,
located at sea level, fires balls at initial speed v = 82 m/s.
(a) At what angle θ from the horizontal must a ball be fired to hit the ship?
(b) What is the maximum range of the cannonballs?

Figure 40: A cannon has attacked a pirate ship


Solution:
(a) We can relate the launch angle θ to the range R by
𝑣02
R= sin 2𝜃
𝑔

1 𝑔𝑅 1 (9.8𝑚⁄ 2 )(560 𝑚)
𝑠
θ = 2 sin-1 𝑣2 = 2 sin-1 (82 𝑚⁄𝑠)2

One solution of sin-1 (54.7°) is displayed by a calculator; we subtract it from 180° to get the other
solution (125.3°). Thus, we get
θ = 27° and θ = 63°
(b) We have seen that maximum range corresponds to an elevation angle θ of 45°.Thus,
𝑣02 (82 𝑚⁄𝑠)2
R= sin 2𝜃 = 2 sin(2 × 45 °)
𝑔 9.8 𝑚⁄𝑠

=686 m
≈ 690 𝑚

55 | P a g e
CHAPTER FIVE: CONCLUSION

In a nutshell, the project can be summarized into the following segments:


i. The concept of Moment of Inertia has been analyzed by the means of solving a number of
problems on Rotational Inertia by general method as well as for complex shaped objects.
ii. While discussing the topic Centre of Gravity, we have come across the definition of
Centre of Gravity along with the general formula for the determination of the Centre of
Gravity. Moreover, differences between the Centre of Gravity and the Centre of mass
have been mentioned.
iii. Virtual Work has been analyzed explicitly by mentioning the work of a force, principle of
Virtual Work, D’ Alembert’s principle etc. Moreover, a number of problems have been
solved as well.
iv. The Equations of Motion have been discussed followed by solving a number of problems
based on those equations.

THANK YOU

56 | P a g e

Você também pode gostar