You are on page 1of 47

Introduction to Mechanics and Mechanisms

Maithripala D. H. S.,
Dept. of Mechanical and Manufacturing Engineering,
Faculty of Engineering,
University of Ruhuna,
Sri Lanka.

August 20, 2007


2
Contents

1 Mechanics 5
1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.2 Particle Motion in Euclidian Space . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.2.1 Kinematics of a Particle Moving in Space . . . . . . . . . . . . . . . . . . . 6
1.2.2 Kinetics of a Particle Moving in Space . . . . . . . . . . . . . . . . . . . . 10
1.3 Relative Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.3.1 Parallely Translating Frames . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.3.2 Rotating Frames . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.3.3 General Moving Frames . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.4 Rigid Body Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
1.4.1 Two-dimensional Rigid Body Rotational Motion . . . . . . . . . . . . . . . 19
1.4.2 Three-dimensional Rigid Body Rotational Motion . . . . . . . . . . . . . . 21
1.4.3 Three-dimensional General Rigid Body Motion . . . . . . . . . . . . . . . 22
1.5 Kinetic Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
1.6 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
1.7 Practice exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

2 Mechanisms 39
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
2.2 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

3
4
Chapter 1

Mechanics

Figure 1.1: Mechanics gone wrong: Tacoma Narrows bridge minutes before collapse.

1.1 Introduction
Historically it is the the study of mechanics that evolved into what we call Engineering today. The
following quote by Wikipedia exemplifies this belief: “Important aspects of the fields of mechanical
engineering, aerospace engineering, civil engineering, structural engineering, materials engineering,

5
6

biomedical engineering and biomechanics were spawned from the study of mechanics”. Mechanics
deals with the scientific description of the world as we perceive it. There are two major branches
of mechanics: classical mechanics and quantum mechanics. The latter has only a history of about
hundred years while the former goes beyond the period of written history. Classical Mechanics is
concerned with describing macroscopic phenomena while quantum mechanics is mainly concerned
with describing microscopic phenomena. It is widely believed that quantum mechanics at “large”
scales approximates classical mechanics. Since at this stage of undergraduate Engineering we are
mostly interested in macroscopic systems we will only consider classical mechanics.
The subject of classical mechanics is generally divided into three branches; rigid-body mechan-
ics, deformable-body mechanics and fluid mechanics. In this class we will concentrate on learning
the basics of rigid-body mechanics while elsewhere in the mechanical Engineering program you
will learn the basics of the other two branches. The study of rigid-body mechanics begins with
describing the motion of a mathematically abstracted point particle that does not occupy any
volume in space. A general rigid body is considered to consist of a large number of point particles
where the distance between each of the particles remain fixed. The geometric description of the
motion of these objects is what is generally known as Kinematics while the study of the cause of
motion is referred to as Kinetics.

1.2 Particle Motion in Euclidian Space


In classical mechanics, to describe the motion of a point particle we need to define and accept three
fundamental concepts; mass, time and space. We associate with each point a quantity called mass
that in a certain sense describes the resistance to change of motion. Time is the quantity that
measures the distance between two separate events and we assume that it has an independent
meaning. We assume that space is a three-dimensional Euclidean space, that is a space where
Euclidean geometry holds.

1.2.1 Kinematics of a Particle Moving in Space


Co-ordinates and Degrees of Freedom
First we will demonstrate how to describe the position of a given point P in a three-dimensional
Euclidean space. Euclidean space is a space where the Euclidean geometric notions such as parallel
lines and normal lines have the usual meaning. Thus to describe a point P in space we begin by
picking a point O in space and setting three mutually perpendicular unit length axis at O. Such
a set of axis is called a ortho-normal frame of reference. Labelling the axis e1 , e2 , e3 to give a right
hand orientation we can symbolically represent the frame as a matrix e = [e1 e2 e3 ]. Using such
a frame any point P in space can be described using three measurements (numbers) x1 , x2 and
Mechanics of Machines: Class notes for ME2204, ME3305, ME4306 7

x3 . These three numbers describe respectively the distance to the point along the e1 , e2 , and e3
axis. Symbolically we describe this as
 
h i x1
 
OP = x1 e1 + x2 e2 + x3 e3 = e 1 e2 e 3  x2  = e x.
x3

The matrix  
x1
x= 
 x2 
x3
is referred to as the representation of the position of the point P and the components are referred
to as the position components of P with respect to the frame e.
If Q is another point in space and if its representation y with respect to e is given by
 
y1
y= 
 y2  ,
y3

then the distance between P and Q is assumed to be given by,


q
d(P, Q) = (x1 − y1 )2 + (x2 − y2 )2 + (x3 − y3 )2 := ||x − y||.

This notion of distance is essentially what it means to be a Euclidean space. It allows us to


mathematically define what a right angle is and also when two straight lines are parallel.
Observe that the introduction of the ortho-normal frame has allowed us to uniquely identify
points in three dimensional Euclidean space with, R3 , the set of ordered triple of numbers. Namely
by identifying each point P with the uniquely corresponding ordered triple (x1 , x2 , x3 ). It has to
be kept in mind that this identification depends heavily on the choice of the ortho-normal frame
e. For instance consider figure 1.2. The position of the particle P is described by the two different
matrices x = [x1 x2 x3 ]T and X = [X1 X2 X3 ]T with reference to the two frames e and b
respectively. Orhto-normal frames are also not the only means of identifying Euclidean space with
R3 . Curvilinear frames can also be used for this identification. For example, as shown in figure
1.3, we may use the quantities (r, θ, φ) to describe the point P . All three sets of ordered triples
(x1 , x2 , x3 ), (X1 , X2 , X3 ) and (r, θ, φ) describe the same point P but note that they are in general
different. Thus the three descriptions provide three different identifications with R3 .
If the point P is moving freely in space (moves without any geometric constraint) then at each
time t, the specification of the three independent measurements is necessary to un-ambiguously
describe the position of the point P at each time t. Since, at each time instant t, three independent
8

Figure 1.2:

quantities (measurements) are needed to describe the position we say that a free particle in space
has three degrees of freedom. In general, a collection of independent measurements (quantities)
required to uniquely describe the position of a moving point particle are called the co-ordinates
of that point and the number of such co-ordinates are called the degrees of freedom (DOF) of
the particle. In the unconstrained situation the three components of the position representation
matrix provide a set of co-ordinates with respect to the frame e. Co-ordinates that describe
the position with respect to an ortho-normal frame are called Euclidean co-ordinates. From the
discussion in the previous paragraph it is clear that Euclidean co-ordinates are not the only type
of co-ordinates. The quantities (r, θ, φ) used to describe the point P are called spherical-polar co-
ordinates. The relationship between various co-ordinates can be found and are called co-ordinate
transformations. For instance the Euclidean co-ordinates of the free particle P are related to the
spherical polar co-ordinates by
   
x1 r sin φ cos θ
x =  x2  =  r sin φ sin θ 
  
.
x3 r cos φ

If the motion of the particle is geometrically constrained so that the quantities (x1 (t), x2 (t), x3 (t))
have to satisfy some scalar expression h(x1 (t), x2 (t), x3 (t)) = 0 then the particle is said to be Holo-
nomically constrained and it can be shown that only two independent measurements are required
to uniquely describe the position of the particle. Thus the constraint has reduced the DOF of the
particle from three to two. If the number of such constraints are two then the DOF of the particle
reduces to one. For instance if the particle is constrained to move on a sphere of radius equal
to one then the constraint equation in the spherical-polar co-ordinates is h(r, θ, φ) = r − 1 = 0
Mechanics of Machines: Class notes for ME2204, ME3305, ME4306 9

Figure 1.3: Description of the point P using Spherical-Polar Co-ordinates.

and the position matrix x will become [sin φ cos θ sin φ sin θ cos φ]T . Therefore the two inde-
pendent measurements (θ, φ) are sufficient to describe the position uniquely and constitutes the
co-ordinates of the constrained P . Since only two measurements (co-ordinates) are required to
uniquely describe the position of P we say that P has two DOF. On the other hand if the particle
is constrained to move in a circle of radius one that lies in the e1 , e2 plane then there are two
constraints, h1 (r, θ, φ) = r − 1 = 0, h2 (r, θ, φ) = φ − π/2 = 0 and the position matrix x will
become [cos θ sin θ 0]T . Therefore the single measurement θ is sufficient to describe the position
uniquely and constitutes the co-ordinate of the constrained P and thus the DOF of P is one. The
general expression
n = 3 − f,
relates the DOF, n, of a particle moving in three dimensional space to the number of holonomic
constraints, f , imposed on that particle. If N points are moving in space then the total degrees
of freedom of the system of points is
n = 3N − f,
where now f is the total number of holonomic constraints of the system of points.
An important fact to keep in mind is that what ever the co-ordinates we use to describe the
point we will always use the matrix x to represent the point in space and that to do so we first
need to specify an ortho-normal frame e. However, the components of this matrix x1 , x2 and x3
will be functions of the co-ordinates.
10

Velocity and Acceleration of a Particle


The velocity, v(t), and acceleration, a, of a point P (t) moving in space is defined with respect
to the ortho-normal frame e with respect to which the position of P (t) is described. That is by
definition the velocity measured in the e frame is the quantity

v(t) = ẋ(t),

and the acceleration measured in the e frame is the quantity

a(t) = v̇(t) = ẍ(t).

Example 1 Consider a particle constrained to move on a sphere of radius one (refer to figure
1.3). The representation of P in an ortho-normal frame e fixed at the center of the sphere is
 
sin φ cos θ
x =  sin φ sin θ 

.
cos φ

Note that the components are a function of the co-ordinates (θ, φ). Then the velocity is
 
φ̇ cos φ cos θ − θ̇ sin φ sin θ
 
v = ẋ =  φ̇ cos φ sin θ + θ̇ sin φ cos θ  .
−φ̇ sin φ

and the acceleration is


 
φ̈ cos φ cos θ − φ̇2 sin φ cos θ − 2φ̇θ̇ cos φ sin θ − θ̇2 sin φ cos θ − θ̈ sin φ sin θ
a = v̇ =  φ̈ cos φ sin θ − φ̇2 sin φ sin θ + 2φ̇θ̇ cos φ cos θ − θ̇2 sin φ sin θ + θ̈ sin φ cos θ 

.
2
−φ̈ sin φ − φ̇ cos φ

1.2.2 Kinetics of a Particle Moving in Space


Kinetics deal with the apparent causes of motion. In this section we investigate the cause of
motion in classical mechanics. Given the present “state” of a particle, the primary interest in
mechanics is to be able to completely predict the future “state” of it. More precisely we need to
know what measurements are needed at the present time in order to be able to uniquely describe
the position of a point of mass m for all future times. For macroscopic motions, the answer to
this deep and profound question was provided by Newton. He postulated that for macroscopic
motions the acceleration a(t) of a given particle was not an independent (arbitrary) quantity and
Mechanics of Machines: Class notes for ME2204, ME3305, ME4306 11

that it was completely specified. More precisely in his second law he stated that in a given frame
e
mẍ = m a(t) = f e (t), (1.1)

where f e (t) was completely known. For macroscopic motion this notion has been verified exper-
imentally. In this expression the mass of the point is taken to be a fundamental property of the
point and is further assumed to be a constant. From a mathematical perspective equation (1.1)
describes a second order differential equation and solving it for x(t) will enable us to predict the
position for all future time. To solve this equation we need to know the right hand side at every
time instant t as well as know the initial conditions x(t0 ) = x0 and v(t0 ) = v0 . Thus knowing the
right hand side turned out to be a crucial step in describing the position and as such was termed
to be the cause of motion. Newton termed this constraint on mass times acceleration, the Force
acting on the particle in the ortho-normal frame e. Thus the force “felt” in a frame e is nothing
but mass times the acceleration in that frame. The force can not be directly measured and can
only be inferred indirectly by resorting to other physical reasoning and justifications.

Example 2 Consider the problem of a horizontal spring with one end fixed to a support and the
other end fixed to point mass P , of mass m. If we give an initial horizontal displacement to the
point mass we know empirically that the point mass will exhibit a simple harmonic motion if the
air and other resistances on the particle is negligible and the motion is small. That is if x(t) is the
displacement of the mass from the unstretched position and if the air resistance on the particle is
negligible and the motion is small the position x(t) of the particle P at a given time t is described
by the second order differential equation

mẍ(t) = −k x(t).

Observing this expression it is evident that the mass times the acceleration of the particle is con-
strained and is equal to −k x(t). Thus we would call −k x(t) the force exerted by the spring on the
particle. Observe that we can not directly measure this instead we can only infer it by measuring
the deflection of the spring.

1.3 Relative Motion


In section 1.2.1 it was pointed out that choosing different frames of reference resulted in different
representations for the position of a given particle. In this section we will see how these different
descriptions with respect to different frames are related to each other.
12

Figure 1.4:

1.3.1 Parallely Translating Frames


Let e be an ortho-normal frame of reference and b(t) be another ortho-normal frame of reference
that is parallely translating with respect to e. We will also refer to b(t) as a moving frame. Let
the origin O0 (t) of the translating frame have the representation o(t) with respect to the frame e.
Consider a point P (t) moving in Euclidian space. This point can be expressed in two different
ways using the two different frames. Let x(t) and X(t) be the representations of the point P (t)
in the e and b(t) frames respectively (refer to figure 1.4). In Euclidian space the meaning of b(t)
being parallel to e is that the two representations x(t) and X(t) of P are related by,

x(t) = o(t) + X(t). (1.2)

Observing that symbolically this means

OP (t) = e x(t) = e o(t) + b(t) X(t)

you will notice that this is what you traditionally wrote down as

OP (t) = OO0 (t) + O0 P (t),

for addition of “vectors”.


Mechanics of Machines: Class notes for ME2204, ME3305, ME4306 13

The velocity of the point P (t) measured in frame e is


v(t) = ẋ(t),
and measured in the moving frame is
V (t) = Ẋ(t).
Differentiating (1.2) we see that the two measured velocities are related by
v(t) = ẋ(t) = ȯ(t) + Ẋ(t) = ȯ(t) + V (t). (1.3)
The acceleration of the point P (t) measured in the e frame is
a(t) = v̇(t) = ẍ(t),
and measured in the moving frame is
A(t) = V̇ (t) = Ẍ(t).
Differentiating (1.3) we see that the two measured accelerations are related by
a(t) = v̇(t) = ö(t) + V̇ (t) = ö(t) + A(t). (1.4)
From (1.4) and (1.1) we have
f e (t) = m a(t) = m(ö(t) + A(t)),
and hence that the total force experienced in the moving frame is
F b (t) = m A(t) = −m ö(t) + f e (t). (1.5)
That is an observer moving with the translating frame will, in addition to the force f e (t) felt in
the e frame, also experience another apparent force
Fapp = −m ö(t). (1.6)
This particular type of apparent forces that arise as a consequence of the translational accelerating
motion of the reference frame are called Einstein forces. Recall that we defined at the beginning
that force felt in a frame is simply the constraint on mass times the acceleration of the particle in
that frame.
Do exercise 11 at this moment. In exercise 11 you are asked to explain why a person standing
on a scale inside an elevator sees his or her weight doubled as the elevator accelerates up at a rate
of g and sees the weight reduced to zero if the elevator decelerates at a rate of g where g is the
gravitational acceleration. You are also asked to show that if, for some reason, the gravitational
force field vanished and that the elevator was moving up at an acceleration of g then the scale
would show the correct weight of the person. This last observation shows that a person inside the
elevator can not distinguish between the following two cases:
14

a.) Gravity is present and the elevator is standing still (or moving at constant velocity).

b.) Gravity is absent and the elevator is accelerating upwards at a rate of g.

It is this observation that led Einstein to the conclusions of General Relativity and in particular
that gravity is an apparent force !!!

1.3.2 Rotating Frames

Figure 1.5:

Let e be an ortho-normal frame of reference and b(t) be an ortho-normal frame of reference


that is rotating with respect to e. We will consider the case where both origins coincide at the
point O in space (rotating frames). Then the two frames are related by

b(t) = e R(t), (1.7)

where R(t) is a special orthogonal matrix (ie RT R = I and det(R) = 1). If we are considering
2-dimensional Euclidian motions then e = [e1 e2 ], b(t) = [b1 (t) b2 (t)] and R(t) is a 2 × 2 matrix
while if it is 3-dimensional Euclidian motions then e = [e1 e2 e3 ], b(t) = [b1 (t) b2 (t) b3 (t)] and
R(t) is a 3 × 3 matrix (refer to figure 1.5). In exercise 8 you are asked to show that the 2 × 2
special orthogonal matrices can be identified with points on the unit circle and hence can be
parameterized by a single angle co-ordinate θ and that the 3 × 3 special orthogonal matrices can
be parameterized using three angle co-ordinates known as Euler angles.
Mechanics of Machines: Class notes for ME2204, ME3305, ME4306 15

The point P (t) can be expressed in both fames as

OP (t) = e x(t) = b(t) X(t) = e R(t)X(t), (1.8)

where x(t) is the representation of the point P (t) with respect to the frame e and X(t) is the rep-
resentation of the point P (t) with respect to the moving frame b(t). Thus the two representations
are related by
x(t) = R(t)X(t). (1.9)
The velocity of the point P (t) measured in the e frame is v(t) = ẋ(t), and measured in the moving
frame b(t) is V (t) = Ẋ(t).
Now differentiating (1.9) we have that
 
v(t) = ẋ(t) = Ṙ(t)X(t) + R(t)Ẋ(t) = R(t) RT (t)Ṙ(t)X(t) + V (t) = R(t)Vb (t). (1.10)

The quantity Vb (t) is given by

Vb (t) = RT (t)Ṙ(t)X(t) + V (t). (1.11)

Observe that Vb is the velocity v(t) expressed using the moving frame b(t). This is only a derived
quantity as compared to the other measured quantities. Since RT (t)R(t) = I, it follows that

ṘT R + RT Ṙ = 0,

and hence that


RT Ṙ = −(RT Ṙ)T = Ω̂,
where Ω̂ is a skew symmetric matrix.
For 2-dimensional Euclidian motion Ω̂ is a 2 × 2 skew-symmetric matrix while it is a 3 × 3
skew-symmetric matrix for 3-dimensional Euclidian motion. In exercise 9 you are asked to show
that 2 × 2 skew-symmetric matrices can be identified with the set of real numbers, R, using the
identification Ω 7→ Ω̂ where " #
0 −Ω
Ω̂ = , (1.12)
Ω 0
and in exercise 10 you are asked to show that the space of 3 × 3 skew-symmetric matrices can
be identified with, the set of ordered triple of real numbers, R3 , using the identification Ω 7→ Ω̂
where  
0 −Ω3 Ω2

Ω̂ =  Ω3 0 −Ω1 , (1.13)
−Ω2 Ω1 0
16

and Ω = [Ω1 Ω2 Ω3 ]T . In 2-dimensional Euclidian motions the quantity Ω corresponds to an


instantaneous rotation of the point P about the fixed point O by an amount equal to Ω. Similarly
it can be shown that in the case of 3-dimensional Euclidian motion the quantity Ω corresponds to
an instantaneous rotation of the point P about the axis Ω through O by an amount equal to the
magnitude ||Ω||. Thus in both cases Ω is defined to be the angular velocity of the point P .
Substituting RT Ṙ = −(RT Ṙ)T = Ω̂ in (1.11) the velocity v expressed using the moving frame
b(t) is given by
RT (t)v(t) = Vb (t) = Ω̂(t)X(t) + V (t). (1.14)
Therefore the two velocities, v(t) and V (t) are explicitly related by
 
v(t) = R(t) Ω̂(t)X(t) + V (t) . (1.15)

The acceleration of the point P (t) measured in the e frame is a(t) = v̇(t) = ẍ(t), and measured
in the moving frame is A(t) = V̇ (t) = Ẍ(t). Differentiating (1.10) we have that
 
a(t) = Ṙ(t)Vb (t) + R(t)V̇b (t) = R(t) RT (t)Ṙ(t)Vb (t) + V̇b (t) ,
 
= R(t) Ω̂(t)Vb (t) + V̇b (t) = R(t)Ab (t), (1.16)

where

Ab (t) = Ω̂(t)Vb (t) + V̇b (t),


˙
= Ω̂2 (t)X(t) + 2 Ω̂(t)V (t) + Ω̂(t)X(t) + A(t). (1.17)

Observe that Ab is the acceleration of the particle in the e frame expressed in the moving frame
b(t). This too is only a derived quantity as compared to the other measured quantities. From
(1.16) and (1.17) the two accelerations, a(t) and A(t) are explicitly related by
 
2 ˙
a(t) = R(t) Ω̂ (t)X(t) + 2 Ω̂(t)V (t) + Ω̂(t)X(t) + A(t) . (1.18)

From (1.17) and (1.1) we have that


 
e 2 ˙
f (t) = m a(t) = m(R(t)Ab (t)) = m R(t) Ω̂ (t)X(t) + 2 Ω̂(t)V (t) + Ω̂(t)X(t) + A(t) , (1.19)

and hence in the rotating moving frame we have that

˙
F b (t) = m A(t) = −m Ω̂2 (t)X(t) − 2m Ω̂(t)V (t) − m Ω̂(t)X(t) + RT f e (t). (1.20)
Mechanics of Machines: Class notes for ME2204, ME3305, ME4306 17

Thus an observer moving with the rotating frame will, in addition to the rotating frame version
of the force felt in the e frame f e (t), feel the effect of the apparent force:

˙
Fapp = −m Ω̂2 (t)X(t) − 2m Ω̂(t)V (t) − m Ω̂(t)X(t). (1.21)

The first term −m Ω̂2 (t)X(t) is known as the Centrifugal force, the second term −2m Ω̂(t)V (t)
˙
is known as the Coriolis force and the third term −m Ω̂(t)X(t) is known as the Euler force.

1.3.3 General Moving Frames


Let e be an ortho-normal frame of reference and b(t) be a moving ortho-normal frame of reference
with origins O and O0 respectively. In this section we consider the case where the frame b(t) is
rotating with respect to the frame e as well as translating with respect to e (ie. the origin O0 of
the b(t) frame is also moving with respect to the frame e). Then a point P (t) can be expressed in
both fames as
OP (t) = e x(t) = e o(t) + b(t) X(t) = e (o(t) + R(t)X(t)) , (1.22)
where x(t) and o(t) are the representations of the points P (t) and O0 (t) in the frame e while X(t)
is the representation of P (t) in the moving frame b(t). Thus the two representations of the point
P are related by
x(t) = o(t) + R(t)X(t). (1.23)
The velocity of the point P (t) measured in the frame e is v(t) = ẋ(t), and measured in the moving
frame is V (t) = Ẋ(t).
Now differentiating (1.23) we have that
 
v(t) = R(t) RT (t)ȯ(t) + Ω̂(t)X(t) + V (t) = R(t)Vb (t), (1.24)

where the quantity Vb (t)


Vb (t) = RT (t)ȯ(t) + Ω̂(t)X(t) + V (t), (1.25)
is the velocity in e expressed with respect to the moving frame b(t). Let Vo (t) = RT (t)ȯ(t) be the
velocity of the point O0 (t) as measured in b(t).
The acceleration of the point P (t) measured in the frame e is a(t) = v̇(t) = ẍ(t), and measured
in the moving frame is A(t) = V̇ (t) = Ẍ(t). Differentiating (1.24) we have that
 
a(t) = Ṙ(t)Vb (t) + R(t)V̇b (t) = R(t) Ω̂(t)Vb (t) + V̇b (t) ,
= R(t)Ab (t), (1.26)
18

where
˙
Ab (t) = V̇o (t) + Ω̂(t)Vo (t) + Ω̂2 (t)X(t) + 2 Ω̂(t)V (t) + Ω̂(t)X(t) + A(t), (1.27)

is the acceleration of the particle in e expressed in the moving frame b(t).


From (1.27) and (1.1) we have that
 
e 2 ˙
f (t) = m a(t) = m R(t) V̇o (t) + Ω̂(t)Vo (t) + Ω̂ (t)X(t) + 2 Ω̂(t)V (t) + Ω̂(t)X(t) + A(t) ,
(1.28)
and hence in the general moving frame we have that
˙
F b (t) = m A(t) = −m V̇o (t) − m Ω̂(t)Vo (t) − m Ω̂2 (t)X(t) − 2m Ω̂(t)V (t) − m Ω̂(t)X(t) + RT f e (t).
(1.29)

1.4 Rigid Body Motion


Consider a rigid body that is free to rotate about a fixed point O. Let e be a “fixed” ortho-normal
frame and let b(t) be a body-fixed ortho-normal frame (which we will call the body frame). Both
frames have coinciding origins at O. The position of a point P in the body, at a time t, is given
by the co-ordinates x(t) in the frame e and by the co-ordinates X in the body frame b(t). Observe
that since the frame b(t) is fixed in the body, the co-ordinates X are independent of time. At
t = 0, the two frames coincide (ie. b(0) = e). We have already seen that the two frames are
related by
b(t) = e R(t)
and hence that
x(t) = R(t)X,
where R(t) is a special orthogonal matrix. It can be shown that in this fashion any given config-
uration of the body can be uniquely identified with a special orthogonal matrix R. Recall that,
since RT Ṙ = Ω̂,
Ṙ = R Ω̂,
where Ω̂ is given by (1.12) for 2-dimensional Euclidian motions and is given by (1.13) for 3-
dimensional Euclidian motions. In 2-dimensional Euclidian motions the quantity Ω corresponds
to an instantaneous rotation of the body about the fixed point by an amount equal to Ω. Similarly
it can be shown that in the case of 3-dimensional Euclidian motion the quantity Ω corresponds
to an instantaneous rotation of the body about the axis Ω by an amount equal to the magnitude
||Ω||. Thus in both cases Ω is defined to be the body angular velocity of the rigid body.
Mechanics of Machines: Class notes for ME2204, ME3305, ME4306 19

Assume that the body is made up of a very large number of point masses. Isolate a partic-
ular point P of mass m and let us analyze the mechanics of this point mass. The velocity and
acceleration of the point P at time t as measured in the inertial frame e is v(t) = R(t)Vb (t) and
a(t) = R(t)Ab (t) respectively where from (1.14)

Vb (t) = Ω̂(t)X, (1.30)

and from (1.17)


˙
Ab (t) = Ω̂2 (t)X + Ω̂(t)X, (1.31)

are their body representations. Let

fpe (t) = fpr (t) + fpext (t)

be the co-ordinates of the total force acting on the point mass P expressed in the inertial frame.
Here fpr is the force that maintains rigidity and arises due to the interaction of the neighboring
particles and fpext (t) is the resultant external force expressed in the inertial frame. Their respective
representations in the body frame are RT (t)fpr (t) = Fpr (t) and RT (t)fpext (t) = Fpext (t). Due to the
assumption of rigidity it follows that Fpr the interaction force as seen in the body is a constant.
Applying Newton’s equations we have that
 
˙
fpe (t) = m a(t) = m R(t)Ab = m R(t) Ω̂2 (t)X + Ω̂(t)X ,
 
˙
RT (t)fpe (t) = m Ω̂2 (t)X + Ω̂(t)X . (1.32)

1.4.1 Two-dimensional Rigid Body Rotational Motion


In this section we only consider the case where the body is fixed at a point O and is free to rotate
about that point in a plane. Chose a body frame b(t) and an inertial frame e such that their
origins coincide at O.
In the case of the two-dimensional Euclidian space, the rotation matrix R is a 2 × 2 special
orthogonal matrix and " #
0 −Ω
Ω̂ = ,
Ω 0
where the real number Ω is defined to be the body angular velocity of the body about O. Thus
using the results shown in exercise 12, Newton’s equation (1.32) in two dimension reduces to
 
˙
m −Ω2 (t)X + Ω̂(t)X = RT (t)fpI (t) = RT (t)(Fpr + Fpext (t)). (1.33)
20

Now if we pre-multiply both sides by X̃ where for X = [ X1 X2 ]T


h i
X̃ = −X2 X1 ,

and sum over all the points in the body we obtain


X  X 
2 ˙
m −Ω (t)X̃X + X̃ Ω̂(t)X = X̃(Fpr + Fpext (t)).

This reduces to X X
Ω̇(t) m ||X||2 = X̃(Fpr + Fpext (t)),
˙
because it can be easily shown that the first term becomes zero and X̃ Ω̂(t)X = Ω̇||X||2 .
Let us individually consider each of these terms. The internal constraint forces that maintain
rigidity are of the nature that they satisfy Newton’t third law, action equals reaction. Thus
X
X̃Fpr = 0.

Let X
X̃Fpext (t) = T ext (t), (1.34)
where X̃Fpext (t) is defined to be the moment of the force Fpext (t) about the axis of rotation and
therefore T ext (t) as the resultant moment of the external forces about the axis of rotation. If we
define X
I= m ||X||2 , (1.35)
as the Moment of Inertia of the 2-dimensional body about the axis of rotation then the Kinetics
of a 2-dimensional rigid body is given by

I Ω̇ = T ext (t).

Thus the complete motion of the rigid body in 2-dimensions is governed by

Ṙ = R Ω̂, (1.36)
I Ω̇ = T ext (t). (1.37)

These equations are referred to as Euler’s 2-dimensional rigid body equations.


Recall that Ω was defined to be the body angular velocity. Consequently the quantity Π = IΩ
is defined to be the body angular momentum. From equation (1.37) it follows that if the external
moment T e = 0 then Π̇ = I Ω̇ = 0. This is known as the law of conservation of body angular
momentum.
Mechanics of Machines: Class notes for ME2204, ME3305, ME4306 21

1.4.2 Three-dimensional Rigid Body Rotational Motion


In this section we consider the case where a point O in the body is fixed and the body is free to
rotate about that point. Chose a body frame b(t) and an inertial frame e such that their origins
coincide at O.
Let us pre-multiply both sides of (1.32) by the 3 × 3 skew symmetric matrix X̂ and sum over
all the points in the body
X  X 
2 ˙
m X̂ Ω̂ (t)X + Ω̂(t)X = X̂(Fpr + Fpext (t)). (1.38)

Consider individually each of the terms in the above equation. The internal constraint forces
that maintain rigidity are of the nature that they satisfy Newton’t third law, action equals reaction.
Thus X X
X̂Fpr = X × Fpr = 0.
Let X X
X̂Fpext (t) = X × Fpext (t) = T ext (t), (1.39)
where X̂Fpext (t) = X × Fpext (t) is defined to be the moment of the force Fpext (t) about the fixed
point of rotation O and therefore T ext (t) as the resultant moment of the external forces about O.
In the exercises you are asked to show that
X ˙
m X̂ Ω̂(t)X = I Ω̇(t),
and X
m X̂ Ω̂2 (t)X = − IΩ(t) × Ω(t),
where X  
I= m ||X||2 I3×3 − XX T , (1.40)
is defined to be the inertia tensor of the body about the fixed point O. Then we have that (1.50)
reduces to
I Ω̇ = IΩ × Ω + T ext .
Thus the complete motion of a rigid body in 3-dimensions is governed by
Ṙ = R Ω̂, (1.41)
I Ω̇ = IΩ × Ω + T ext . (1.42)
Recall that Ω(t) was defined to be the body angular velocity. The quantity Π(t) = I Ω(t) is defined
to be the body angular momentum. Using properties of cross products and dot products it can be
easily shown that
IΩ ∙ (IΩ × Ω) = 0.
22

Consequently if the resultant external moment of the forces is zero, T ext = 0, we have that

d d√ IΩ ∙ I Ω̇
||Π(t)|| = IΩ ∙ IΩ = √ = 0.
dt dt IΩ ∙ IΩ
This implies that if the resultant external moment about the fixed point is zero then the magnitude
of the body angular momentum is a constant. This is known as the law of conservation of
momentum for a rigid body.
The inertia tensor (27) is a symmetric matrix and it can be shown that it is positive definite.
Consequently it is always possible to find a body frame such that the Inertia tensor is diagonalized.
In particular if the the body is axi-symmetric and the body frame is aligned along the axis of
symmetry then  
I1 0 0
I=  0 I2 0  ,

(1.43)
0 0 I3
and (1.42) takes the form

I1 Ω̇1 = (I2 − I3 )Ω2 Ω3 + T1e , (1.44)


I2 Ω̇2 = (I3 − I1 )Ω3 Ω1 + T2e , (1.45)
I3 Ω̇3 = (I1 − I2 )Ω1 Ω2 + T3e , (1.46)

where now I1 , I2 , I3 are called the principle moments of inertia.

1.4.3 Three-dimensional General Rigid Body Motion


In this section we consider the case where the body describes a general Euclidean motion. Chose a
body frame b(t) and an inertial frame e. From (1.29) a point P fixed in the moving frame satisfies
the following relationship.
˙
m V̇o (t) + m Ω̂(t)Vo (t) + m Ω̂2 (t)X(t) + m Ω̂(t)X(t) = Fpr + Fpext (t). (1.47)

First let us sum (1.47) overall the points in the body.


X  X  X X X X
˙
m V̇o (t)+ m Ω̂(t)Vo (t)+ Ω̂2 (t) m X(t)+ Ω̂(t) m X(t) = Fpr + Fpext (t). (1.48)

Now if we choose the b(t) frame so that its origin O coincides with the center of mass of the rigid
body, then the last two terms on the left hand side of (1.48) become zero and we have

M V̇o (t) + M Ω̂(t)Vo (t) = F ext (t), (1.49)


Mechanics of Machines: Class notes for ME2204, ME3305, ME4306 23
P P P
where M = m, F ext (t) = Fpext (t) and we have set Fpr = because the constraint forces occur
in equal and opposite pairs.
Let us now pre-multiply both sides of (1.47) by the 3 × 3 skew symmetric matrix X̂ and sum
over all the points in the body
X X X ˙ X X X
m X̂ V̇o (t) + m X̂ Ω̂(t)Vo (t) + m X̂ Ω̂2 (t)X(t) +
m X̂ Ω̂(t)X(t) = X̂Fpr + X̂Fpext (t).
(1.50)
If the body frame b(t) is fixed at the center of mass of the body then the fist two terms become
zero and we end up with the rigid body equations (1.42) for a purely rotational rigid body derived
in section 1.4.2, Thus the complete equations of motion for general rigid body motion where the
body frame b(t) is fixed at the center of mass of the body are

Ṙ(t) = R(t)Ω̂(t), (1.51)


ȯ(t) = R(t)V0 (t), (1.52)
M V̇o (t) = −M Ω̂(t)Vo (t) + F ext (t), (1.53)
I Ω̇ = IΩ × Ω + T ext . (1.54)

1.5 Kinetic Energy


The kinetic energy KEp of a particle of mass m is defined with respect to an inertial frame and
is defined to be
1
KEp = m||v(t)||2 , (1.55)
2
where || ∙ || is the Euclidian norm in R3 and is defined to be ||X||2 = X T X = X ∙ X (physically
this gives the length of the vector X).
Since ||RX|| = ||X|| with respect to a frame b(t) rotating purely with respect to the inertial
frame we also have that
1 1 1
KEp = m||v(t)||2 = m||RT (t)v(t)||2 = m||Vb (t)||2 .
2 2 2
Hence from (1.14) we have that,
1 1  
KEp = m||Vb (t)||2 = m −X T Ω̂2 X + 2V T Ω̂X + ||V ||2 . (1.56)
2 2
If the particle P is fixed with respect to the purely rotating frame then V (t) = Ẋ(t) = 0 and
hence
1 1  
KEp = m||Vb (t)||2 = m −X T Ω̂2 X . (1.57)
2 2
24

To compute the total energy of a purely rotating rigid body we sum up the kinetic energy of
all the particles to obtain
X 1X 1X  
KE = Tp = m||Vb (t)||2 = m −X T Ω̂2 X . (1.58)
2 2
In 2-dimensions it is easy to show that
X 1 1
−m X T Ω̂2 X = IΩ2 ,
2 2
and hence that
1
KE = IΩ2 . (1.59)
2
In exercise 27 you are asked to show that in 3-dimensions
X 1 1
−m X T Ω̂2 X = ΩT IΩ,
2 2
and hence that
1
KE = ΩT IΩ. (1.60)
2
On the other hand with respect to a general rotating and translating body where the body
frame is fixed at the center of mass of the body we can show that the kinetic energy is given by
M 1
KE = ||V0 ||2 + ΩT IΩ. (1.61)
2 2
Mechanics of Machines: Class notes for ME2204, ME3305, ME4306 25

1.6 Exercises
Exercise 1 A bead is constrained to move on a frictionless wire that lies on a horizontal plane.
The wire is bent to a shape of a parabola (ie. y = x2 ). Find the constraint forces that keep the
bead on the wire and the equation of motion of the bead.

Exercise 2 Two point masses P1 and P2 are moving in three-dimensional Euclidean space such
that P1 moves on a sphere and the distance between P1 and P2 remain fixed. What is the degrees of
freedom of this system of particles and how many co-ordinates would you need to uniquely describe
the motion of the system. Also specify the configuration space and a suitable set of coordinates for
the system.

Exercise 3 In a Euclidean frame e, with no external force fields present, a particle P is con-
strained to move such that ||x||2 = xT x = constant, where x is the representation of the particle
in the e frame (observe that this corresponds to a particle motion on a sphere if the Euclidean space
is three-dimensional while it corresponds to a particle motion on a circle if the Euclidean space
is two-dimensional). Show that the velocity of the particle is always orthogonal to the position x,
ie. xT ẋ = 0. Differentiating this expression show that the motion of the particle in the e frame is
described by
1 ||ẋ||2
ẍ = fce (t) = − x,
m ||x||2
where fce (t) is the representation of the constraint force in the e frame.
Let b(t) be a frame such that its origin coincides with that of e and the particle P appears to
be fixed with respect to b(t). Let b(t) = e R(t) and x(t) = R(t)X where X is the representation of
the point P in b(t) and is a constant.
T 2 x T (RT Ω̂2 R)x
Ω̂ X
1. Show that the constraint force is fce (t) = m X||X||2 RX = m ||x||2
x, where Ω is the
angular velocity of the particle.

2.

Exercise 4 In a 2D-Euclidean frame e, with no external force fields present, a particle is con-
strained to move on a circular wire with no friction. If x is the representation of the particle in
the e frame show that the motion of the particle is described by

ẍ = −Ω2 x,

where Ω is the angular velocity of the particle. How would you relate this to the results of exercise
6.
26

Exercise 5 In a 2D-Euclidean frame e, with no external force fields present, a particle is con-
strained to move on a circular wire with no friction. If x is the representation of the particle in
the e frame show that the motion of the particle is described by
ẍ = −Ω2 x,
where Ω is the angular velocity of the particle. How would you relate this to the results of exercise
6.

Exercise 6 A bead is constrained to move on a frictionless wire that lies on a horizontal plane.
The wire is bent to a shape of a circle. Picking suitable co-ordinates find the constraint forces that
keep the bead on the circle and the equation of motion of the bead.

Exercise 7 A cannon is fired with an initial velocity of 1 ms and a firing angle of 45o . Find the
horizontal distance to the point of landing of the cannon.

Exercise 8 Show that 2 × 2 special orthogonal matrices R can be identified with points on the
unit circle S 1 and hence that they can be parameterized using the single angle co-ordinate θ. Also
show how to parameterize 3 × 3 special orthogonal matrices R using Euler angles.

Exercise 9 Show that the space of 2 × 2 skew-symmetric matrices can be identified with R using
the identification Ω 7→ Ω̂ where " #
0 −Ω
Ω̂ = . (1.62)
Ω 0
Using this identification show that Ω̂2 = −Ω2 I2×2 where I2×2 is the 2 × 2 identity matrix.

Exercise 10 The space of 3 × 3 skew-symmetric matrices can be identified with R3 using the
identification Ω 7→ Ω̂ where  
0 −Ω3 Ω2
Ω̂ = 
 Ω3 0 −Ω1  , (1.63)
−Ω2 Ω1 0
and Ω = [Ω1 Ω2 Ω3 ]T . Using this identification show that
Ω̂X = Ω × X.

Exercise 11 — Einstein’s box experiment: Explain why a person standing on a scale inside an
elevator sees his or her weight doubled as the elevator accelerates up at a rate of g and sees the
weight reduced to zero if the elevator decelerates at a rate of g. Also show that if, for some reason,
the gravitational force field vanished and the elevator was moving up at an acceleration of g then
the scale would still show the correct weight of the person.
Mechanics of Machines: Class notes for ME2204, ME3305, ME4306 27

Exercise 12 — Particle motion in 2-dimensions: Consider a particle P moving in 2-dimensional


Euclidian space. Using the results of section 1.3.2 describe the inertial velocity and inertial accel-
eration of the particle P as expressed in the rotating ortho-normal frame b(t) = [b1 (t) b2 (t)] (refer
to figure 1.6). The frame e = [e1 e2 ] is an ortho-normal inertial frame. Write down the apparent
forces acting on a particle as observed in the moving frame.

Figure 1.6:

Exercise 13 — Motion of a particle fixed in a rotating frame: If a particle P appears to be fixed


in the moving frame of exercise 12 describe the inertial velocity and inertial acceleration of the
particle P as observed in the rotating frame. What are the constraint forces required to keep the
particle fixed in the rotating frame. If the rate of rotation of b(t) is a constant in the counter
clockwise direction, θ̇(t) = ψ, describe the motion of the particle P in the inertial frame e and
show that the motion corresponds to a constant rate counter clockwise circular motion. Verify
your conclusions using MATLAB simulations.

Exercise 14 — Observing a fixed particle in a rotating frame: A particle P is fixed in an ortho-


normal inertial frame, e, at a distance l from the origin. Show that an observer moving with a
rotating ortho-normal frame, with origin O coinciding with that of e, rotating at a constant rate of
ψ in the counter clockwise direction will see it rotating clockwise in a circle at a constant angular
rate of ψ. Verify your answer using MATLAB simulations. Show that the observer in the moving
frame will think that a constant radial force equal to lψ 2 is acting on the particle.

Exercise 15 — 2-dimensional particle motion in polar co-ordinates: Using the results of Exercise
12, for a particle moving in 2-dimensions, write down explicitly the inertial velocity and inertial
acceleration as expressed in the moving frame b(t) = [er (t) eθ (t)] (refer to figure 1.7). Additionally
if θ̇(t) = ψ, a constant, then write down the inertial velocity and inertial acceleration of the particle
28

Figure 1.7:

Figure 1.8:

as expressed in the moving frame b(t) = [er (t) eθ (t)]. In each case write down the apparent forces
acting on a particle as observed by an observer in the moving frame.

Exercise 16 A point P (t) appears to be fixed in the moving frame b(t) = [b1 (t) b2 (t) b3 (t)] write
down the forces acting on the particle as observed by an observer in the moving frame.

Exercise 17 Consider a ball of mass m constrained to move as shown in figure 1.8. The disk is
rotating at a constant angular rate of Ω. Write down the equations of motion of the ball and also
the constraint forces.

Exercise 18 Consider a ball of mass m constrained to move as shown in figure 1.9. The Disk is
rotating at a constant angular rate of Ω. Write down the equations of motion of the ball.
Mechanics of Machines: Class notes for ME2204, ME3305, ME4306 29

Figure 1.9:

Exercise 19 Consider the bead on a rotating hoop shown in figure 1.10. The hoop is rotating
about the vertical axis in the counter clockwise direction at a constant angular rate of Ω. Neglecting
friction between the bead and the hoop write down the governing equations that describe the motion
of the bead in a co-ordinate frame fixed on the hoop. Also write down the constraint forces acting
on the bead.

Exercise 20 Consider the systems in figure 1.11. The shaft with the arm rotates about the vertical
axis in the counter clockwise direction. The spring is pivoted to the arm so that it can freely rotate
in the plane of the shaft and the arm. Write down the equations that describe the motion of the
point P in a co-ordinate frame that is fixed on the shaft. The mass of point P is m.

Exercise 21 For the system shown in figure 1.12,


(i.) Derive the equations of motion as observed by an observer in an inertial frame.
(ii.) Derive the equations of motion as observed by an observer moving with the box (ie.
in the moving frame co-ordinates).
30

Figure 1.10:

Exercise 22 For each of the systems shown in figures 1.13 to 1.18, derive the governing differ-
ential equations using Newton’s law.

Exercise 23 Consider the ball inside the swinging hoop shown in figure 1.19. The hoop has a
mass distribution of ρ per unit length and a radius of r. Neglecting friction between the ball and the
hoop write down the constraint forces acting on the ball. Write down Euler’s rigid body equations
for the hoop and using this expression eliminate the constraint forces appearing in the equations
for the ball and find the equations that govern the motion of the entire system.

Exercise 24 — Simulation and Experimentation: Consider the case of two observers E and B.
Observer E is positioned in a 2-dimensional ortho-normal inertial frame, e = [e1 e2 ] and the
observer B is positioned in a rotating ortho-normal frame b(t) = [b1 (t) b2 (t)]. Both frames have
coinciding origins. Discuss the following for the case where the frame b(t) with observer B is
rotating at a constant angular rate of θ̇(t) = ψ in the counter clockwise direction.

a.) Observer B takes a particle of unit mass and slowly releases it on the rotating plane along the
b1 (t) axis at a unit distance away from the origin. Neglecting friction between the plane and
the particle, simulate using MATLAB, the motion of the particle as seen by B.

b.) Observer B takes a particle of unit mass and constraints it so that it can move only along the
b1 (t) axis. He then slowly releases it at time t = 0 at a unit distance away from the origin.
Find the constraint force at a time t.
Mechanics of Machines: Class notes for ME2204, ME3305, ME4306 31

Figure 1.11:

c.) However observer B is aware that no external force fields are present. Therefore having learnt
Newtonian mechanics in his undergraduate years B realizes that the frame in which he is
making the measurements is rotating. Device and carry out an experiment the observer B
can use to determine if the frame is rotating at a constant speed or not and find the rate of
rotation if the answer is yes. (Hint: You can use a setup similar to that of the Centrifugal
force apparatus in the Applied Mechanics lab.)

d.) Describe using MATLAB the motion of the particle in part a.) as seen by the observer E.

e.) Assume that a central inward force field is acting in the inertial frame where the intensity
is equal to r12 Newtons per unit mass. A particle of unit mass is released in the rotating
frame at a distance of 0.5 units away from the origin and with zero velocity. Using results
of exercise 12 predict the behavior of the particle as observed in the rotating frame and using
MATLAB simulate the motion of the particle in the rotating frame.

f.) Explain using the results of e.) why a Hurricane that has formed in the Northern hemisphere
has a counter clockwise rotation and a Hurricane that has formed in the Southern hemisphere
has a clockwise rotation (see figure 1.20).

Exercise 25 — Self study on vibrating MEMS Gyroscopes: The operation of a vibrating MEMS
Gyroscope relies on the presence Coriolis forces. Write a two page report on vibrating MEMS
gyroscopes. Include in the report an explanation of its working principle, its use, and methods of
fabrication.
32

Figure 1.12:

Figure 1.13:

Exercise 26 Show that x(t) × fpe (t) = R(t) (X × Fpe (t)). (Hint: First show using the properties
of cross product that R(a × b) = Ra × Rb)
Exercise 27 Using the following properties of cross products in R3
1. A × (B × C) = (A ∙ C)B − (B ∙ A)C
2. A ∙ (B × C) = C ∙ (A × B) = B ∙ (C × A)
3. (Ω × X) ∙ (ω × X) = (Ω ∙ ω)||X||2 − (Ω ∙ X)(ω ∙ X)
show that
X 1 1
−m X T Ω̂2 X = ΩT IΩ,
2 2
where X  
I= m ||X||2 I3×3 − XX T ,
Mechanics of Machines: Class notes for ME2204, ME3305, ME4306 33

Figure 1.14:

Figure 1.15:

is defined to be the inertia tensor of the body about the fixed point O, and
X ˙
m X̂ Ω̂(t)X = I Ω̇,

and X
m X̂ Ω̂2 (t)X = − IΩ × Ω.

Exercise 28 — Gimbal Gyroscope: Consider the Gyroscope shown in figure 1.21. Show using the
rigid body equations for an axi-symmetric body, (1.44) — (1.46), that a.) if no external torques
are applied on the system the spin axis direction will remain a constant. b.) an external torque
applied on the frame about the z axis will induce a motion about the x axis. Verify your answers
using the Gyroscopic experimental setup in the applied mechanics lab.

1.7 Practice exercises


1. For the single bar pendulum shown in figure 1.22 derive the equations of motion using Euler’s
2-dimensional rigid body equations and also write down the kinetic energy of the system.
34

Figure 1.16:

Figure 1.17:

2. For the double bar pendulum shown in figure 1.23 write down the kinetic energy of the
system.
Mechanics of Machines: Class notes for ME2204, ME3305, ME4306 35

Figure 1.18:

Figure 1.19:

(a) A Southern Hurricane (Catarina). (b) A Northern Polar Hurricane.


Figure 1.20: Figures show the clockwise and anti-clockwise rotation of respectively a southern
hemisphere and northern hemisphere formed hurricane. Figures are courtesy of Wikipedia.
36

Figure 1.21: The Gimbal Gyroscope

Figure 1.22:
Mechanics of Machines: Class notes for ME2204, ME3305, ME4306 37

Figure 1.23:
38
Chapter 2

Mechanisms

2.1 Introduction
A coupling of rigid bodies through various types of connections is called a kinematic chain. The
various types of connections are called kinematic pairs while each constituent rigid body is called
a link. A kinematic chain is said to be open if at least one link is coupled to the others only
through one kinematic pair and is said to be closed otherwise. Kinematic pairs are classified by
their degree of movability s (ie. the degrees of freedom of the relative motion of the two links
joined by the pair). Revolute, prismatic and screw are examples of s = 1 while cylindric and
roll-slide pairs are examples of s = 2. Higher pairs of movability are spheric and planar — s = 3,
cylinder-plane — s = 4, sphere-plane — s = 5. The degrees of freedom (DOF), W , of a kinematic
chain is given by
5
X
f = 6N − (6 − s)ns , (2.1)
s=0

where N is the total number of links and ns is the total number of kinematic pairs with movability
s. When the chain is constrained to move in a two-dimensional Euclidian plane then (2.1) reduces
to
2
X
f = 3N − (3 − s)ns . (2.2)
s=0

A kinematic chain is said to be a structure if the couplings are such that no mobility remains
in the chain. A kinematic chain is said to be a mechanism if at least one link is “fixed”. A
mechanism is capable of transmitting force and/or motion from one place to another. The fixed
link of a mechanism is usually referred to as the frame and the motion of the rest of the links
are expressed with respect to this link. By fixing different links of a given chain we can obtain
different mechanisms. This is called kinematic inversion. It can be seen that the relative motion

39
40

of all kinematic inversions are identical. The number of DOF of a mechanism is also given by
(2.1) and (2.2) with N being replaced by N − 1 to account for the loss of DOF due to fixing one
link.
The first step in the design or analysis of a mechanism is to sketch an equivalent skeleton or
kinematic diagram. Conventionally in this diagram the links are numbered and the joints are
lettered with the fixed link always numbered as 1. The next step is to represent the mechanism
using a graph whose vertices correspond to links and whose edges correspond to kinematic pairs
where the number of edges joining two nodes correspond to the number of DOF, ns , of the
kinematic pair that joins the two links represented by the nodes. These graphs enable one to
easily calculate the number of DOF and furthermore allows the use of graph theory to design and
analyze mechanisms. The next step in the design and analysis process is the problem of geometric
analysis.
If a given mechanism has n DOF then there exists n number of independent quantities q =
(q1 , q2 , ∙ ∙ ∙ , qn ) called configuration co-ordinates that uniquely describe the configuration of the
mechanism. The output of a mechanism are a m number of functions yk = φ(q) k = 1, 2, ∙ ∙ ∙ , m
of the configuration co-ordinates q = (q1 , q2 , ∙ ∙ ∙ , qn ). It is only rarely that one can find these
functions in explicit form. Typically one finds m number of constraint equations between the qk ’s
and the yk ’s that need to be solved using numerical methods. Sometimes the inverse problem of
finding the qk for a given yk is the primary concern. The next step in the design and analysis is
to obtain the corresponding relationships between the instantaneous velocities. The last step is to
find the relationship between the generalized input forces and the generalized output forces. This
is obtained by neglecting the dynamics of the mechanism and considering energy conservation.
For two bodies in relative motion it can be shown that there exists a common point with
zero instantaneous velocity. Such a point is called an instantaneous center. An important result
applicable for two-dimensional mechanisms is what is known as Kennedy’s theorem. This states
that three rigid bodies in relative motion with each other will have their respective instantaneous
centers lying in a straight line.

x = Y + R(φ)X, (2.3)
ˆ = 0,
ẋ = Ẏ + R(φ)φ̇X (2.4)
ˆ
φ̇X = −R(−φ)Ẏ , (2.5)

X = φ̇R(−φ)Ẏ (2.6)
φ̇2
Proof of Kennedy’s theorem

x = R(θ)X1 = Y + R(φ)X2 , (2.7)


Mechanics of Machines: Class notes for ME2204, ME3305, ME4306 41

ẋ = R(θ)ˆθ̇X1 = R(φ)φ̇X
ˆ ,
2 (2.8)
ˆθ̇R(θ)X = φ̇R(φ)X
ˆ , (2.9)
1 2
ˆ
ˆθ̇ (Y + R(φ)X ) = φ̇R(φ)X
2 2, (2.10)
θ̇
R(φ)X2 = Y, (2.11)
(φ̇ − θ̇)
φ̇
x = Y, (2.12)
(φ̇ − θ̇)

2.2 Exercises
1. Obtain the corresponding graph of the mechanisms shown in figures figure 2.1 — figure 2.8
and find their number of DOF.

2. For the slider crank mechanism shown in figure 2.4

Figure 2.1: Digger Mechanism


42

Figure 2.2: Aircraft Door Mechanism

Figure 2.3: Hood Mechanism


Mechanics of Machines: Class notes for ME2204, ME3305, ME4306 43

Figure 2.4: Windshield Wiper Mechanism

Figure 2.5: Windshield Wiper Mechanism


44

Figure 2.6: Crank and Rocker Mechanism

Figure 2.7: Cam and Follower Mechanism


Mechanics of Machines: Class notes for ME2204, ME3305, ME4306 45

Figure 2.8: Plier Mechanism


46
Bibliography

[1] Haim Baruh, Analytical Dynamics McGraw Hill, 1999.

[2] R. Abraham and J. E. Marsden, Foundations of Mechanics, Second Ed. Westview, 1978.

[3] V. I. Arnold, Mathematical Methods of Classical Mechanics, Second Ed., Springer-Verlag, New
York 1989.

[4] J. E. Marsden and T. S. Ratiu, Introduction to Mechanics and Symmetry, Second Ed. Springer-
Verlag, New York 1999.

[5] F. Bullo and A. D. Lewis, Geometric Control of Mechanical Systems: Modeling, Analysis, and
Design for Simple Mechanical Control Systems, Springer-Verlag, New York 2004.

47

You might also like