You are on page 1of 107

2103-617 Advanced Dynamics

Thitima Jintanawan
Preface

I write this book, shortly because I love to do it. With this book, I would like to
share my own experience on Dynamics from education and researches for over
ten years with students and the people in the same field. The book is organized
and written from my viewpoints of dynamics, and it is appropriate for the one
who study dynamics in the intermediate level. Why written in English? It is about
the right time and right situation. When Chulalongkorn University started to
promote the faculties to carry out the class in English in the academic year of
2001, I joined the program and have an opportunity to teach the Advanced
Dynamics class in English. I started to write the first draft in English for the
class's lecture notes and continuingly improve it since then. It is the right situation
when we have the ME graduate foreign student attending the class in the Year
2005. Language is not the obstacle for communication. Instead good writing
communication needs a well-organized manuscript that indeed my book still has a
room for improvement.

My first experience in dynamics during the undergraduate years is not different


from everyone's experience in that we simply start with the Newton's 2nd Law, and
laws of energy and momentum. Taking the motion of a particle as an example,
both the Newton's Law and the principle of energy in dynamics have the same
root from the law of linear momentum. Later, when I was doing my Master
Degree, I learnt the whole new aspect of dynamics, namely, 3-D Dynamics,
Dynamic Model and Analysis, Derivation of Equations of Motion, Lagrange's
Mechanics, Stability and Rotordynamics. During the time for PhD, I got the first
lesson of dynamics there from my advisor. Not as a coursework requirement, he
kindly gave me the intensive lectures on dynamics and vibration of deformable
bodies such as plate and shell, so that I could have a necessary background to start
the research. Next I began to learn the Halmiton’s Principle and the Variational
Principle from several courseworks and from self-study. These two principles are
fundamentals of Lagrange's mechanics. Truly, My PhD research is the best lesson
of dynamics that I have learnt. At that time, I started to use Matlab as a program
tool for dynamic simulation and continue writing the Matlab codes nowadays.
Also, one of a good memory for Dynamics during those years is the opportunity
to attend the seminar "a New Paradigm of Dynamics" by a world famous
dynamist who develops the Kane’s method, Prof. Thomas Kane of Stanford
University.

There are two premium dynamists who are my role model. The first person is my
supervisor at the University of Melbourne, Dr. Januzt Krodkiewski. He is an icon
of the discipline and logic. The second one is my PhD advisor, Prof. Steve Shen.
He is an icon of making things simple (no matter how complicate they are). Both
of them similarly have an excellent background in Mathematics. I wish I could be
a half of the people that I admire. Therefore, it is to them that I dedicate this book.

Thitima Jintanawan
2

Copyright
c 2005
Thitima Jintanawan
Contents

1 Kinematics 5
1.1 Evolution of Kinematics . . . . . . . . . . . . . . . . . . . . . 5
1.2 Position Vector, Velocity, and Acceleration . . . . . . . . . . . 5
1.3 Angular Velocity . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.4 Rate of Change of a Constant-Length Vector . . . . . . . . . 8
1.5 Moving Coordinate Systems . . . . . . . . . . . . . . . . . . . 10
1.6 Coordinate Transformation . . . . . . . . . . . . . . . . . . . 17
1.6.1 First set of Euler angles–precession-nutation-spin (φθψ) 20
1.6.2 Second set of Euler angles–yaw-pitch-row (ψθφ) . . . . 20
1.7 Angular velocity related to Euler angles . . . . . . . . . . . . 23
1.8 A Finite Motion . . . . . . . . . . . . . . . . . . . . . . . . . 25
1.8.1 Transformation matrices for a finite rotation . . . . . 26
1.8.2 Transformation matrices for a general motion . . . . . 30

2 Linear and Angular Momentums 35


2.1 Dynamics of a System of Particles: a Review . . . . . . . . . 35
2.1.1 Total mass . . . . . . . . . . . . . . . . . . . . . . . . 35
2.1.2 First moment of mass . . . . . . . . . . . . . . . . . . 35
2.1.3 Linear momentum . . . . . . . . . . . . . . . . . . . . 37
2.1.4 Angular momentum . . . . . . . . . . . . . . . . . . . 37
2.1.5 Moment of force . . . . . . . . . . . . . . . . . . . . . 38
2.1.6 Laws of linear and angular momentum . . . . . . . . . 38
2.2 Angular Momentum of a Rigid Body . . . . . . . . . . . . . . 40
2.3 Mass Moment of Inertia . . . . . . . . . . . . . . . . . . . . . 42

3 Dynamics of a Rigid Body 45


3.1 Newton-Euler Equations of a rigid body . . . . . . . . . . . . 45
3.2 Modified Euler’s equations . . . . . . . . . . . . . . . . . . . . 49
3.3 Introduction to stability of a spin body . . . . . . . . . . . . 53

3
4 CONTENTS

4 Multi-Body Mechanical System 57


4.1 Degrees of Freedom (DOF) . . . . . . . . . . . . . . . . . . . 57
4.2 Constraints . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
4.3 Constraint Equations . . . . . . . . . . . . . . . . . . . . . . . 57
4.4 Classification of Constraints . . . . . . . . . . . . . . . . . . . 60
4.5 Number of DOF vs. Driving Forces . . . . . . . . . . . . . . . 61
4.6 Dynamic Analysis of Multi-Body Mechanical Systems . . . . 61
4.7 Example Problem: Dynamics of Two-Link Arms . . . . . . . 62

5 Principle of Virtual work 69


5.1 Virtual Displacement and Virtual Work . . . . . . . . . . . . 69
5.2 Holonomic and Nonholonomic Constraints . . . . . . . . . . . 69
5.3 Generalized Coordinates and Jacobian . . . . . . . . . . . . . 71
5.4 Principle of Virtual Work . . . . . . . . . . . . . . . . . . . . 75
5.5 D’Alembert principle . . . . . . . . . . . . . . . . . . . . . . . 76

6 Lagrange Mechanics 81
6.1 Kinetic Energy . . . . . . . . . . . . . . . . . . . . . . . . . . 81
6.2 Potential Energy . . . . . . . . . . . . . . . . . . . . . . . . . 82
6.3 Remarks on Properties of Generalized Coordinates for the
System with Holonomic constraints . . . . . . . . . . . . . . . 84
6.4 Derivation of Lagrange’s equations . . . . . . . . . . . . . . . 85
6.5 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
6.6 Lagrange Multiplier . . . . . . . . . . . . . . . . . . . . . . . 94

7 Stability Analysis 99
7.1 Equilibrium, Quasi-Equilibrium, and Steady States . . . . . . 99
7.2 Stability of Equilibrium or Steady State . . . . . . . . . . . . 99
Chapter 1

Kinematics

In this chapter various coordinate systems, such as cartesian and cylindri-


cal coordinates, are introduced. Position vector, velocity and acceleration
of particles and rigid bodies are formulated using different reference coordi-
nates. Each coordinate system is related to the other through the coordinate
transformation. For the 3-D transformation, two different sets of Euler an-
gles: precession-nutation-spin and yaw-pitch-roll, are conventionally used.
Finally the transformation matrix used to describe a finite motion of rigid
bodies is revealed.

1.1 Evolution of Kinematics


Prior to 1950s: Express velocity v and acceleration a in terms of scalar
components and use graphical method to determine total magnitude
and direction

1950s and later: Express velocity v and acceleration a using vector ap-
proach

Recent years: Express the rotation with a matrix and utilize the matrix
operation for calculating the cross product. The matrix approach can
be simply implement in a computer simulation program.

1.2 Position Vector, Velocity, and Acceleration


Fig. 1.1 shows a particle moving in a 3-dimensional (3-D) space. Let’s
introduce a cartesian or rectangular coordinate system XY Z as shown in

5
6 CHAPTER 1. KINEMATICS

rz
particle

r
moving path

k
j ry
i Y
O
rx

Figure 1.1: A cartesian coordinate system

Fig. 1.1 in which all coordinates are orthogonal to each other and its axes
do not change in direction. If we choose XY Z in Fig. 1.1 as an inertial or
fixed reference frame1 , the absolute motion of the particle in Fig. 1.1 can be
described by a position vector r as follows
r = rx i + ry j + rz k (1.1)
where i, j, and k are the unit vectors of XY Z and rx , ry , and rz are scalar
components of r in X, Y , and Z coordinates. The position vector r can be
alternatively presented in a matrix form as a 3 × 1 column matrix given by
r = [ rx ry rz ]T (1.2)
Note that the position vector r must be measured from the origin O of the
chosen inertial frame.
Figure 1.2 shows another set of coordinate system so called the cylindri-
cal coordinates ρθz, with their unit vectors eρ eθ ez . In Fig. 1.2, the position
vector r expressed in terms of eρ eθ ez is
r = ρeρ + zez (1.3)
The absolute velocity v is defined as a time derivative of the position vector
r given by
dr
v = (1.4)
dt
= ṙx i + ṙy j + ṙz k
= ρ̇eρ + ρθ̇eθ + żez
1
An inertial or fixed reference frame is the coordinate system whose origin O is fixed
in space
1.3. ANGULAR VELOCITY 7

z
Z
ez

θ
ρ

ρ
r

Y
θ

Figure 1.2: A cylindrical coordinate system

The absolute acceleration a is defined as a time derivative of the velocity v


given by

dv
a = (1.5)
dt
= r̈x i + r̈y j + r̈z k
   
= ρ̈ − ρθ̇ 2 eρ + ρθ̈ + 2ρ̇θ̇ eθ + z̈ez

1.3 Angular Velocity


Figure 1.3 shows a rigid cylinder having a rotation about n axis. The abso-
lute angular velocity ω of the rigid body is defined as


ω = n (1.6)
dt
= ω 1 e1 + ω 2 e2 + ω 3 e3

where ω1 , ω2 , and ω3 are components of the angular velocity in an arbitrary


rectangular coordinate system with unit vectors e1 , e2 , and e3 . The angular
velocity can be expressed in a matrix form as
 
0 −ω3 ω2
 
ω̃ =  ω3 0 −ω1  (1.7)
−ω2 ω1 0
8 CHAPTER 1. KINEMATICS

e3

v e2
r
A e1
θ(t)

Figure 1.3: Angular velocity

The velocity at point A in Fig. 1.3 is then

v = ω×r (1.8)
≡ ω̃r

Equation (1.9) indicates that the cross product can be represented by the
matrix multiplication or
    
v1 0 −ω3 ω2 r1
    
v =  v2  =  ω3 0 −ω1   r2  (1.9)
v3 −ω2 ω1 0 r3

Note that for the matrix multiplication in (1.9), components of ω̃ and r


must be expressed in the same coordinate system.

1.4 Rate of Change of a Constant-Length Vector


The Theorem in the Vector of Calculus states that “The time derivative of
a fixed length vector c is given by the cross product of its rotation rate ω
and the vector c itself.”
dc
=ω×c (1.10)
dt
Example 1.1 :
The defense jet plane as shown in Fig. 1.4 operates in a roll maneuver with
rate of φ̇ and simultaneously possesses a yaw maneuver (turn to left) with
a rate of ψ̇. Determine a relative velocity of point C on the horizontal
stabilizer at coordinates (b, a, 0), observed from the C.G. of the plane.
1.4. RATE OF CHANGE OF A CONSTANT-LENGTH VECTOR 9

X ez

ex G

C
ey

Figure 1.4: A defense jet plane

Solution
From Fig. 1.4, the position vector of point C relative to the C.G. is a
fixed length vector given in terms of the body coordinate system as
r(rel)
c = bex + aey = [ b a 0 ]T (1.11)
Hence the relative velocity of C is
vc(rel) = ω × r(rel)
c ≡ ω̃r(rel)
c (1.12)
where ω = φ̇ey + ψ̇ez is the rotation rate or the angular velocity of the
reference coordinates moving with the body. ω can be written in a matrix
form as  
0 −ψ̇ φ̇
 
ω̃ =  ψ̇ 0 0 
−φ̇ 0 0
Therefore
  
0 −ψ̇ φ̇ b
  
vc(rel) =  ψ̇ 0 0    = [ −aψ̇ bψ̇ −bφ̇]T
a
−φ̇ 0 0 0
As another example, the unit vectors ijk for any rotating system of
coordinates xyz is also the fixed length vector. Hence the rate of change of
these ijk vectors can be determined from the same theorem as i̇ = ω × i,
j̇ = ω × j, and k̇ = ω × k, where ω is the angular velocity of such rotating
coordinate system xyz.
10 CHAPTER 1. KINEMATICS

o
y

path of origin o
x

Figure 1.5: Translating coordinate systems

z Z

o Y

X
x

Figure 1.6: Rotating coordinate systems

1.5 Moving Coordinate Systems


Any moving coordinate system xyz used to describe the motion can be
divided into 3 types depending on its motion with respect to the inertial
frame XY Z. They are

1. Translating coordinate systems (Fig. 1.5)

2. Rotating coordinate systems (Fig. 1.6)

3. Translating and rotating coordinate systems (Fig. 1.7)

A moving coordinate system, chosen such that it is attached to a moving


body, is normally used as a reference frame to describe kinematics of the
1.5. MOVING COORDINATE SYSTEMS 11

z
Z

path of origin o

x
Y

Figure 1.7: Translating and rotating coordinate systems

body. Specifically, such reference coordinate system is arranged such that


its origin o is fixed to and translate with the body’s C.G. and its axes
synchronously rotate with the body.
Fig. 1.8 shows a particle moving in 3-D space. XY Z is an inertial frame
with unit vectors ijk. Also xyz is the moving reference frame with unit
vectors e1 e2 e3 . If the angular velocity of xyz is ω and the position vector r
is
r =R+ρ (1.13)

Then the velocity v of the particle is given by

dr
v = (1.14)
dt
dR dρ
= +
dt dt
= Ṙ + vr + ω × ρ

In (1.14), Ṙ = dR
dt is the velocity of the origin o of the reference frame
 
xyz

and ω is its angular velocity. In addition vr , sometimes denoted by dt ,
rel
is a relative velocity of the particle with respect to xyz or the relative ve-
locity observed by the observer moving (both translating and rotating) with
xyz, whereas ddtρ is the relative velocity observed by the observer who only
translates but not rotates with xyz. ω × ρ in (1.14) is hence the difference
between these two relative velocities. If ρ is

ρ = ρ1 e1 + ρ2 e2 + ρ3 e3 (1.15)
12 CHAPTER 1. KINEMATICS

ω
z
e3
Z path of moving
O' e2 y reference frame
e1
x
ρ

R path of particle
r
k
j
i Y
O

Figure 1.8: Moving coordinate systems

Then  

vr ≡ = ρ̇1 e1 + ρ̇2 e2 + ρ̇3 e3 (1.16)
dt rel

The absolute acceleration a of the particle is then

dv
a = (1.17)
dt
dρ dω
= R̈ + ar + ω × vr + ω × + ×ρ
dt dt
From (1.14)

= vr + ω × ρ (1.18)
dt
Plug (1.18) into (1.17) yields

a = R̈ + ar + ω × ω × ρ + ω̇ × ρ + 2ω × vr (1.19)

where ar = ρ̈1 e1 + ρ̈2 e2 + ρ̈3 e3 .


We can describe the physical meaning of each term in (1.19) as follows.

• R̈ is the acceleration of the origin o of the moving reference frame xyz.

• ar is the relative acceleration of the particle as observed in the moving


reference frame xyz.
1.5. MOVING COORDINATE SYSTEMS 13

• ω × ω × ρ is a centripetal acceleration, or the correction term for the


local position vector ρ considering that the observer rotates with the
moving reference frame.

• ω̇ × ρ is another correction term for the angular acceleration vector ω̇


of the moving reference frame.

• 2ω × vr is the Coriolis acceleration which is the other correction term


from two sources, both of which measure the rotation of the basis
(unit) vectors of the moving reference frame and associate with an
interaction of motion along more than one coordinate curve.
Example 1.2 :
The satellite shown in Fig. 1.9 has a steady spin Ω about the body fixed axis
e3 The solar panel arm rotates about the e2 -axis with a rate θ̇, and angular
acceleration θ̈ = 0. The panel arm also moving along the radial direction er
with a steady rate ṡ = α. Determine an absolute acceleration of the point
P at the end of the solar panel.
Solution:
Let [e1 e2 e3 ] be the coordinate system that rotates with the body. Hence
ω e1 e2 e3 = Ωe3 . Choose [er eθ e2 ] in Fig. 1.9 as a rotating reference frame.
For this case we obtain the terms in (1.13) and (1.14) as

R = be1 , ρ = ρr er + ρθ eθ + ρ2 e2 = (s(t) + c)er

ω = Ωe3 + θ̇e2
Note that R and ω are the fixed-length vectors with constant magnitudes.
From (1.19), the absolute acceleration of point P is

ap = R̈ + ar + ω × ω × ρ + ω̇ × ρ + 2ω × vr (1.20)

where

R = be1 , Ṙ = Ωe3 × R = bΩe2 , R̈ = Ωe3 × Ṙ = −bΩ2 e1

vr = ρ̇r er + ρ̇θ eθ + ρ̇2 e2 = ṡ(t)er = αer


ar = ρ̈r er + ρ̈θ eθ + ρ̈2 e2 = s̈(t)er = 0
ω̇ = Ωe3 × ω = −Ωθ̇e1
Components in (1.20) are now expressed in terms of two different coordinate
systems [e1 e2 e3 ] and [er eθ e2 ]. To express these terms in only one coordinate
system, i.e. [e1 e2 e3 ], we need the coordinate transformation.
14 CHAPTER 1. KINEMATICS

e3

er
P

θ
s(t)
+c eθ
P
θe2

e1

e2
b

Figure 1.9: A satellite

er e3

ur
u θ u3
e1 u1 e2
θ uθ

Figure 1.10: Coordinate systems [e1 e2 e3 ] and [er eθ e2 ]


1.5. MOVING COORDINATE SYSTEMS 15

From Fig. 1.10, we obtain the transformation relation of an arbitrary


vector u as
      
ur cosθ sinθ u3 u3
u= = =T
uθ −sinθ cosθ u1 u1

The reader can prove that T is an orthogonal matrix or T−1 = TT . As


a result
   −1      
u3 cosθ sinθ ur −1 ur T ur
= =T =T
u1 −sinθ cosθ uθ uθ uθ

With the coordinate transformation, we can express all terms in (1.20) in


terms of [e1 e2 e3 ] components as
vr = αer = α (cosθe3 + sinθe1 ) = [ αsinθ, 0, αcosθ ]T

ω = [ 0 θ̇ Ω ]T
 
0 −Ω θ̇
 
ω̃ =  Ω 0 0 
−θ̇ 0 0
ω̇ = [ −Ωθ̇ 0 0 ]T
 
0 0 0
˙  
ω̃ =  0 0 Ωθ̇ 
0 −Ωθ̇ 0
ρ = (s(t)+c)er = (s(t)+c) (cosθe3 + sinθe1 ) = (s(t)+c)[ sinθ, 0, cosθ ]T
R̈ = [ −bΩ2 0 0 ]T
Plug these terms into (1.20), we obtain ap in terms of the rotating system
of coordinates [e1 e2 e3 ] as
ap = [ −bΩ2 0 0 ]T + 0   
0 −Ω θ̇ 0 −Ω θ̇ sinθ
   
+(s(t) + c)  Ω 0 0  Ω 0 0  0 
 −θ̇ 0 0  −θ̇ 0 0 cosθ
0 0 0 sinθ
   (1.21)
+(s(t) + c)  0 0 Ωθ̇   0 
 0 −Ωθ̇  0 
cosθ
0 −Ω θ̇ sinθ
  
+2α  Ω 0 0  0 
−θ̇ 0 0 cosθ
16 CHAPTER 1. KINEMATICS

z1
z2
.
α

L
G y2
µ

β

y1

Figure 1.11: A ventilator

Now your task is to follow the previous procedure and express ap in terms
of the coordinate system [er eθ e2 ].
Example 1.3 :
Figure 1.11 shows a ventilator mounted on a rotating base. The base has
an oscillatory motion with α = α0 sinωt. The rotor spins with a constant
angular velocity Ω in the direction shown. Its center of gravity G is offset
by µ from the axis of rotation. Determine:
1. components of the absolute angular velocity of the rotor along the
system of coordinates fixed to the rotor.
2. components of the absolute velocity of the center of gravity of the rotor
along the same system of coordinates.
Solution:
Figure 1.12 shows the coordinate systems XY Z, x1 y1 z1 , x2 y2 z2 , and
xyz. From Fig. 1.12 coordinate transformations are:
    
y2 cosβ sinβ y1
=
z2 −sinβ cosβ z1
    
x cosαt −sinαt x2
=
z sinαt cosαt z2
1.6. COORDINATE TRANSFORMATION 17

Z, z1 y1
α
z2
Y
α
z

Ωt
X
x1

z1 y2, y
z2

Ωt
x2
y2
β
x
β
x1, x2 y1

Figure 1.12: Coordinate Systems

Absolute angular velocity of the rotor is then

ω = α̇ez1 + Ωey2
= α̇(sinβey2 + cosβez2 ) + Ωey2
= (α̇sinβ + Ω)ey2 + α̇cosβez2
= (α̇sinβ + Ω)Ωey + α̇cosβ(−sinΩtex + cosΩtez )
= −α̇cosβsinΩtex + (Ω + α̇sinβ)ey + α̇cosβcosΩtez

The position vector of G is

rG = Ley2 + µez
= Ley + µez
= [ 0 L µ ]T

Absolute velocity of G is

ṙG = ω× rG


0
 
= ω̃  L 
µ

1.6 Coordinate Transformation


In Section 1.5, we simply transform the coordinates in two dimensional
(2-D) space. Now let’s consider a general 3-D coordinate transformation.
Specifically, we want to establish a transformation matrix C that transform
components of a vector in one system of coordinates to another system of
18 CHAPTER 1. KINEMATICS
Z
z

y
r

iI
X
x

Figure 1.13: A vector r and two sets of coordinate systems XY Z and xyz

coordinates. Let XY Z be an inertial reference frame with unit vectors IJK,


and xyz be a rotating coordinate system with unit vectors ijk as shown in
Fig. 1.13. An arbitrary vector r in Fig. 1.13 can be expressed as

r = rX I + rY J + rZ K
= rx i + ry j + rz k (1.22)

Components of r in XY Z coordinates are

rX = r · I = (rx i + ry j + rz k) · I
rY = r · J = (rx i + ry j + rz k) · J
rZ = r · K = (rx i + ry j + rz k) · K (1.23)

Or

rX = rx cos iI + ry cos jI + rz cos kI


rY = rx cos iJ + ry cos jJ + rz cos kJ
rZ = rx cos iK + ry cos jK + rz cos kK (1.24)

(1.24) can be put in a matrix form as


   
rX rx
   
 rY  = C  ry  (1.25)
rZ rz
1.6. COORDINATE TRANSFORMATION 19

Ji

iJ

iY J

Figure 1.14: components of the unit vectors

where C is a matrix of directional cosines so called a coordinate transfor-


mation matrix given by
 
cos iI cos jI cos kI
 
C =  cos iJ cos jJ cos kJ  (1.26)
cos iK cos jK cos kK

Note that C is the orthogonal matrix where C−1 = CT . This yields


   
rx rX
  T  
 ry  = C  rY  (1.27)
rz rZ

Q: Are all nine components of C independent?


To answer this question, let’s consider Fig. 1.14, showing the following
relations.
|iY |
cos iJ = = |iY | (1.28)
|i|
With similar expressions for the other axes, we come up with the following
6 relationships

|iX |2 + |iY |2 + |iZ |2 = cos2  iI + cos2  iJ + cos2  iK = 1


|jX |2 + |jY |2 + |jZ |2 = cos2  jI + cos2  jJ + cos2  jK = 1
|kX |2 + |kY |2 + |kZ |2 = cos2  kI + cos2  kJ + cos2  kK = 1
|Ix |2 + |Iy |2 + |Iz |2 = cos2  Ii + cos2  Ij + cos2  Ik = 1
|Jx |2 + |Jy |2 + |Jz |2 = cos2  Ji + cos2  Jj + cos2  Jk = 1
|Kx |2 + |Ky |2 + |Kz |2 = cos2  Ki + cos2  Kj + cos2  Kk = 1

With these 6 relations of the directional cosines, there are only 9 − 6 = 3


independent components of C. Specifically, only three independent angular
20 CHAPTER 1. KINEMATICS

transformation terms are needed to describe the coordinate transformation.


There exist many possible sets of angular transformation, but two popular
sets called Euler angles are normally used. Each set consists of three angles
describing the sequence of rotations as described in the following subsections.

1.6.1 First set of Euler angles–precession-nutation-spin (φθψ)


This set of Euler angles is normally used to describe the gyroscopic systems
such as rotordynamics. The sequence of rotations as shown in Fig. 1.15 is

• Precession: rotation about Z axis by φ(t) to get x y  z  or x y  Z

• Nutation: rotation about x axis by θ(t) to get x y  z  or x y  z 

• Spin: rotation about z  axis by ψ(t) to get xyz or xyz 

The coordinate transformations are then


      
rX cosφ −sinφ 0 rx rx
      
 rY  =  sinφ cosφ 0   ry  = C1  ry  (1.29)
rZ 0 0 1 rz  rz 
      
rx 1 0 0 rx rx
      
 y  
r  = 0 cosθ −sinθ  y 
r  = C 2  ry   (1.30)
rz  0 sinθ cosθ rz  rz 
      
rx cosψ −sinψ 0 rx rx
      
 ry  =  sinψ cosψ 0   ry  = C3  ry  (1.31)
rz  0 0 1 rz rz
Combine (1.29)-(1.31), therefore
   
rX rx
   
 rY  = C1 C2 C3  ry  (1.32)
rZ rz

1.6.2 Second set of Euler angles–yaw-pitch-row (ψθφ)


This set of Euler angles is normally used to describe the dynamics of vehicles.
The sequence of rotations as shown in Fig. 1.16 is

• Yaw: rotation about Z axis by ψ(t) to get x y  z  or x y  Z


1.6. COORDINATE TRANSFORMATION 21

Z z’

φ
y’

φ
Y

φ
X
x’

z’
z’’

θ y’’

θ y’

x’ x’’

z’’ z
ψ
y
y’’
ψ

x’’
x

Figure 1.15: First set of Euler’s angles and sequence of rotation


22 CHAPTER 1. KINEMATICS

Z z’
ψ
y’
ψ
Y
Z X ψ
x’
ψ z’
z’’
θ
θ

θ y’ y’’
φ θ
Y x’
x’’
X
z’’ z
y
φ φ y’’

x’’ x φ

Figure 1.16: Yaw, pitch, and roll axes of vehicle dynamics


1.7. ANGULAR VELOCITY RELATED TO EULER ANGLES 23

• Pitch: rotation about y  axis by θ(t) to get x y  z  or x y  z 

• Roll: rotation about x axis by φ(t) to get xyz or x yz

The coordinate transformations are then


      
rX cosψ −sinψ 0 rx rx
      
 rY  =  sinψ cosψ 0   ry  = [Rψ ]  ry  (1.33)
rZ 0 0 1 rz  rz 
      
rx cosθ 0 sinθ rx rx
      
 ry  =  0 1 0   ry  = [Rθ ]  ry  (1.34)
rz  −sinθ 0 cosθ rz  rz 
      
rx 1 0 0 rx rx
      
 ry  =  0 cosφ −sinφ   ry  = [Rφ ]  ry  (1.35)
rz  0 sinφ cosφ rz rz
Therefore    
rX rx
   
 rY  = [Rψ ] [Rθ ] [Rφ ]  ry  (1.36)
rZ rz

1.7 Angular velocity related to Euler angles


For the first set of Euler angles, the absolute angular velocity ω of xyz
coordinates is given by

ω = φ̇k + θ̇ex + ψ̇ez


≡ ω x ex + ω y ey + ω z ez (1.37)

Rewrite ex and k in terms of ex , ey and ez , using (1.29)-(1.31), we get


    
ωx sinθsinψ cosψ 0 φ̇
    
 ωy  =  sinθcosψ −sinφ 0   θ̇  (1.38)
ωz cosθ 0 1 ψ̇

Similarly, for the second set of Euler angles, the absolute angular velocity
ω of xyz coordinates is given by

ω = ψ̇k + θ̇ey + φ̇ex


= ω x ex + ω y ey + ω z ez (1.39)
24 CHAPTER 1. KINEMATICS

φ
x

θ y’

Figure 1.17: Yaw, pitch, and roll axes of vehicle dynamics

Rewrite ey and k in terms of ex , ey and ez , using (1.33)-(1.35), we get


    
ωx 1 0 −sinθ φ̇
    
 y  
ω = 0 cosφ cosθsinφ   θ̇  (1.40)
ωz 0 −sinφ cosθcosφ ψ̇

(1.38) and (1.40) relate the Euler angles, the rotation that measured in real
applications, with the components of the angular velocity, ωx , ωy and ωz , in
the reference coordinate system.
Example 1.4 :
A submarine shown in Fig. 1.17 undergoes a yaw rate ψ̇ = AcosΩt and a
pitch rate θ̇ = BsinΩt. If the local x-axis is in the long-body direction,
describe the velocity of the bow of the submarine relative to its center of
mass.
Solution:
From Fig. 1.17, let xyz with their unit vectors ex , ey , and ez be the body-
fixed rotating system of coordinates. The velocity of the bow observed from
the submarine C.G. is then
v =ω×ρ (1.41)

where
ρ = Lex (1.42)

and ω is the angular velocity of the body or the angular velocity of the xyz
coordinates given by
ω = ψ̇k + θ̇ey (1.43)
1.8. A FINITE MOTION 25

φ
A'

Figure 1.18: A finite motion of a rigid body

ω in terms of ex , ey , and ez can be obtained from (1.40) as


    
ωx 1 0 −sinθ 0
    
ω =  ωy  =  0 cosφ cosθsinθ   θ̇  (1.44)
ωz 0 −sinφ cosθcosφ ψ̇
Note that, in this case, the submarine performs only pitch and yaw rotations
but no row. Neglecting the higher order terms, ω is therefore
ω = −ψ̇sinθex + θ̇ey + ψ̇cosθez (1.45)
Substitution of (1.42) and (1.45) into (1.41) yields
v = Lψ̇cosθey − Lθ̇ez (1.46)
From given ψ̇ = AcosΩt and θ̇ = BsinΩt, and if ψ(0) = θ(0) = 0, then
ψ(t) = AΩ sinΩt and θ(t) = − Ω cosΩt. Substitution of these conditions into
B

(1.46) yields
 
B
v = −ALcosΩtcos cosΩt ey − LBsinΩtez (1.47)

 
Ω , cos Ω cosΩt ≈ 1. In addition if A = B, the
For the small value of B B

velocity vector v = −AL(cos Ωtey + sin Ωtez ) performs a circular path.

1.8 A Finite Motion


A general motion of any rigid body can be resolved into the translation u
of an arbitrary point on the body and a finite rotation φ about this point
26 CHAPTER 1. KINEMATICS

as shown in Figure 1.18. First we consider the transformation matrix for a


finite rotation. Then the transformation matrix for a general finite motion,
possessing both translation and rotation, is considered.

1.8.1 Transformation matrices for a finite rotation


Define a position vector of any point P on the body before and after the
rotation as rp and rp , respectively. A transformation matrix T relating rp
and rp is given by
rp = Trp (1.48)

Properties of the transformation matrix T are described as follows:

1. Because of no deformation of a rigid body, T is the same for any point


p in the body. Hence the subscript p in (1.48) can be drop out.

r = Tr (1.49)

2. The rotation should be invertible.

r = T−1 r (1.50)

3. The length of r is unchanged, hence


 T 
r · r = rT r = r · r = r r (1.51)

Or
  T 
rT r = r r (1.52)
= (Tr)T Tr
= rT TT Tr

Hence
TT T = I (1.53)

(1.53) indicates that T−1 = TT or T is the orthogonal matrix.

The transformation T for a rotation about Z-, Y -, and X-axes can be de-
termined subsequently as follows.
1.8. A FINITE MOTION 27

(x', y', z')

X
(x, y, z)

Figure 1.19: Finite rotation about Z-axis

1. Rotation about Z-axis with φ


If the previous coordinates of any point p is r = [ x y z ]T and
the new coordinates of this point is r = [ x y  z  ]T as seen in
Figure 1.19, then
      
x cosφ −sinφ 0 x x
       
 y  =  sinφ cosφ 0   y  = T1  y  (1.54)
z 0 0 1 z z
The transformation matrix T1 in this case is
 
cosφ −sinφ 0
 
T1 =  sinφ cosφ 0  (1.55)
0 0 1

2. Rotation about Y -axis with θ


From Figure 1.20 we obtain the transformation matrix T2 for a rota-
tion about Y -axis as
 
cosθ 0 sinθ
 
T2 =  0 1 0  (1.56)
−sinθ 0 cosθ

3. Rotation about X-axis with ψ


From Figure 1.21, the transformation matrix T3 for a rotation about
X-axis is  
1 0 0
 
T3 =  0 cosψ −sinψ  (1.57)
0 sinψ cosψ
28 CHAPTER 1. KINEMATICS

X
θ

Figure 1.20: Finite rotation about Y -axis

Figure 1.21: Finite rotation about X-axis


1.8. A FINITE MOTION 29

θ
L
φ b

Figure 1.22: A falling box

In general the transformation matrices do not commute; i.e., T1 T2 = T2 T1 .


However, for the infinitesimal angular displacement, cosφ ≈ 1, sinφ ≈ φ and
so on. Also φ ≈ ωz ∆t, θ ≈ ωy ∆t, and ψ ≈ ωx ∆t. In this case T1 , T2 , and
T3 do commute. If the rigid body has an infinitesimal rotation about an
arbitrary axis during time ∆t, the new position vector r related to the
previous position vector r, according to (1.48), is then

r = r(t + ∆t) = T1 T2 T3 r(t) (1.58)

Substitution of (1.55), (1.56), and (1.57) into (1.58) yields


  
1 −ωz ∆t ωy ∆t x(t)
  
r(t + ∆t) =  ωz ∆t 1 −ωx ∆t   y(t)  (1.59)
−ωy ∆t ωx ∆t 1 z(t)

Therefore the velocity v is given by


  
0 −ωz ωy x(t)
r(t + ∆t) − r(t)   
v = lim =  ωz 0 −ωx   y(t)  ≡ ω̃r (1.60)
∆t→0 ∆t
−ωy ωx 0 z(t)

From (1.60), it is proved that the angular velocity ω̃ can be represented


in a matrix form as previously introduced in (1.7).
The transformation matrix for a finite rotation is useful for a com-
puter graphic programming simulating the dynamics of rigid-body motion
as shown in the following example.
Example 1.5 :
30 CHAPTER 1. KINEMATICS

θ
L
φ b
mg
O
R .
Cθ a

Figure 1.23: A free body diagram of the falling box

A box, considered as the planar problem, is hinged as shown in Figure 1.22.


Construct the Matlab m-file to simulate the dynamics of this falling box.
Solution:
First we need to derive the equation governing the motion of this box. A
free body diagram (FBD) of the box is shown in Figure 1.23. The dynamics
of the falling box is governed by the law of angular momentum given by

[ Mo = Ḣo ]; mgLsin(θ + φ0 ) − C θ̇ = Io θ̈ (1.61)

where C is the torsional damping coefficient used to model the friction at


the hinge and Io is the mass moment of inertia about o. To solve (1.61),
let’s define state variables as x1 = θ and x2 = θ̇ Then (1.61) can be written
in state form as
   
ẋ1 x2
= (1.62)
ẋ2 esin(x1 + φ0 ) − cx2

where e = mgL C
Io and c = Io . (1.62) together with the transformation matrix
for the finite rotation in (1.55) are used in the MatLab program to determine
the new position of the falling box. The detail of this program is presented
in Figure 1.24 and the result is shown in Figure 1.25.

1.8.2 Transformation matrices for a general motion


A general motion of a rigid body as shown in Fig. 1.26 can be divided into
two parts: a translation u and a finite rotation θ. The position vector r
1.8. A FINITE MOTION 31

clear all
% MATLAB Animation Program for Falling Box
%===Define the vertices of the box
a=0.1;
b=0.2;
x=[0 a a 0 0];
y=[0 0 b b 0];
%===Define a matrix whose column vectors are the box vertices
r=[x; y];
%===Draw the box in the initial position
figure(1), clf
axis([-0.3 0.3 -0.3 0.3])
line(x, y,'linestyle','--');
grid on
%===Define parameters
m=1;
g=9.81;
C=0.001;
L=0.5*sqrt(a^2+b^2);
I=m*(a^2+b^2)/12;
e=m*g*L/I;
c=C/I;
%===Define initial conditions
theta = 0;
omega = 0;
phi_0 = atan(a/b);
%===steps
dt = 0.001; % time step for simulation
n=10; % # of animation
M=moviein(n); % # define a matrix M for movie in
%========= Finish data input ============================

%===Numerically integrate the equations of motion using Newton method


for j = 1:n; % Do loop for new box graphic

for n =1:20; % Do loop for elapsed time integration


omega = omega+dt*e*sin(theta+phi_0)-dt*c*omega;
theta=theta+dt*omega;
end
%===Rotate box graphic using finite rotation matrix
A=[cos(theta) sin(theta); -sin(theta) cos(theta)];
r1=A*r;
x1=r1(1,:);
y1=r1(2,:);
patch(x1,y1,'r');
axis('equal')
M(:,j)=getframe;
end
%===Show movie
%figure(2), clf
%movie(M,1,2);

Figure 1.24: Matlab program for animation of box falling


32 CHAPTER 1. KINEMATICS

0.25

0.2

0.15

0.1

0.05

-0.05

-0.1

-0.15

-0.2

-0.25
-0.2 -0.15 -0.1 -0.05 0 0.05 0.1 0.15 0.2 0.25 0.3

Figure 1.25: Simulation of the falling box

ez

A

r
ex
u
k
A

j y
i θ

Figure 1.26: Finite motion


1.8. A FINITE MOTION 33

describing the finite rotation of the rigid body is then


r = u + ρ (1.63)
= u + Aρ
where A is the transformation matrix of the rotation relating ρ and ρ . In
addition r = [ x y z ]T and u = [ ux uy uz ]T expressed in the inertial
reference coordinates xyz, and ρ = [ ρx ρy ρz ]T expressed in the local
coordinate system ex ey ez . Substituting the component vectors into (1.64),
the finite motion in Fig. 1.26 is governed by
      
x ux cosθ −sinθ 0 ρx
      
 y  =  uy  +  sinθ cosθ 0   ρy  (1.64)
z uz 0 0 1 ρz

If r and ρ are expanded as r = [ x y z 1 ]T and ρ = [ ρx ρy ρz 1 ]T ,


(1.64) can be rewritten as
r4×1 = T4×4 ρ4×1 (1.65)
where T is the transformation matrix for a general finite motion given by
 
| ux
 | 
 A3×3 uy 
 
T= | uz 
 
 −− −− −− −|− −− 
0 0 0 | 1
For a finite translation, the transformation matrix is simply
 
1 0 0 | ux
 0 1 0 | uy 
 
 
T1 =  0 0 1 | uz 
 
 −− −− −− −|− −− 
0 0 0 | 1
For a finite rotation, the transformation matrix is simply
 
| 0
 A3×3 | 0 
 
 
T2 =  | 0 
 
 −− −− −− −|− −− 
0 0 0 | 1
Note that T = T1 T2 .
34 CHAPTER 1. KINEMATICS
Chapter 2

Linear and Angular


Momentums

This chapter is organized into three parts: 1) dynamics of a system of par-


ticles; 2) an angular momentum of a rigid body; and 3) a mass moment of
inertia. These topics are used as fundamentals for a study of dynamics of a
rigid body and a multi-body mechanical system in Chapter 3 and Chapter
4, respectively.

2.1 Dynamics of a System of Particles: a Review


Figure 2.1 shows a system consisting of n-particles in 3-D where the i-th
particle is subjected to the applied force Fi . Also c is the center of mass or
center of gravity (C.G.) of the system.

2.1.1 Total mass


A total mass of the system shown in Figure 2.1 is

M= mi (2.1)

where mi is a mass of the i-th particle and is the sum over i, for i =
1, 2, . . . , n.

2.1.2 First moment of mass


The first moment of the total mass about its C.G. is the sum of the first
moment of each mass given by:

rc M = ri mi (2.2)

35
36 CHAPTER 2. LINEAR AND ANGULAR MOMENTUMS

Fi
mi

vi
mN
FN
Z ρi
ri

F1 m
rc
c

m1 Y
o
F3
X

m3
F2
m2

Figure 2.1: A system of particle

From (2.2), the center of mass rc can be obtained as

1 
rc = ri mi (2.3)
M

From Figure 2.1, the displacement of the i-th particle relative to the C.G. is

ρi = ri − rc (2.4)

In addition, sum of the first moment of each mass about C.G. is given by

 
ρ i mi = [r − rc ] mi
 i 
= ri mi − rc mi
(2.5)
= rc M − rc M
= 0

Equation (2.5) indicates that sum of the first moment of each mass about
the system’s C.G. is zero.
2.1. DYNAMICS OF A SYSTEM OF PARTICLES: A REVIEW 37

2.1.3 Linear momentum


A linear momentum P of the system of particles is defined as follows:

P ≡ mi vi
 dri
= mi
dt
d  
= mi ri (2.6)
dt
d
= (M rc )
dt
= M vc

2.1.4 Angular momentum


An angular momentum is defined as the first moment of the linear momen-
tum. The angular momentum of the system of particles about the origin o
is then given by 
Ho ≡ ri × mi vi (2.7)
The angular momentum of the system of particles about the system’s C.G.
is defined as 
Hc ≡ ρi × mi ρ̇i (2.8)
Ho and Hc are related through the following equation

Ho = Hc + rc × M vc (2.9)

To prove the relation (2.9), we rewrite (2.7) as follows



Ho = [(ρi + rc ) × mi vi ]
  (2.10)
= ρi × mi vi + rc × mi vi

The second term on the right of (2.10) is then



rc × mi vi = rc × M vc (2.11)

The first term on the right of (2.10) can be rewritten as


 
ρi × mi vi = ρ × mi (ṙc + ρ̇i )
 i 
= ρ × mi ṙc + ρi × mi ρ̇i
 i (2.12)
= (mi ρi ) × ṙc + Hc
= 0 + Hc

Substitution of (2.11) and (2.12) into (2.10), therefore, yields (2.9).


38 CHAPTER 2. LINEAR AND ANGULAR MOMENTUMS

2.1.5 Moment of force


The moment due to all applied forces about o is

Mo = ri × Fi (2.13)

2.1.6 Laws of linear and angular momentum


The laws of linear and angular momentum relate the applied forces and
moments to the linear and angular momentums of the system.

Law of linear momentum


Applying the Newton’s 2nd law to each i-th particle, we obtain

mi v̇i = Fi + fij ; i = j (2.14)
j

where Fi are external forces applied to the mass mi , and fij is a reaction
force that the j-th particle acts on the the i-th particle. Also note that
fij = −fji . Summation of (2.14) for all particles then yields
  
mi v̇i = Fi + fij ; i = j (2.15)
i i i j


Since fij = −fji , therefore fij = 0. Hence (2.15) becomes
i j

dP 
= M v̇c = Fi (2.16)
dt i

(2.16) is the law of linear momentum for the system of particles, stating that
the rate of change of linear momentum of the system is equal to the sum of
all external forces applied to the system.

Law of angular momentum


Let’s take the first moment of (2.14) about o and sum over all particles:
 
   
(ri × mi v̇i ) = (ri × Fi ) + ri × fij  ; i = j (2.17)
i i i j
2.1. DYNAMICS OF A SYSTEM OF PARTICLES: A REVIEW 39

Now consider (2.17) term by term. The term on the left is rewritten as
  dvi
(ri × mi v̇i ) = ri × mi
dt
i i  
d  (2.18)
= ri × mi vi
dt i
= Ḣo

The first term on the right of (2.17) is (ri × Fi ) ≡ Mo , and the second
i
term on the right is zero as shown in the following proof.
Let’s consider any two particles m and n. Since fmn = −fnm and rm − rn
is approximately colinear with fmn , then the action-reaction pair of any
arbitrary internal moments are zero, or

rm × fmn + rn × fnm = (rm − rn ) × fmn = 0 (2.19)

According to (2.19), therefore


 
 
ri × fij  = 0 (2.20)
i j

Substituting (2.18) to (2.20) into (2.17), we get


dHo
= Ḣo = Mo (2.21)
dt
Equation (2.21) is the law of angular momentum, stating that the rate of
change of angular momentum about o is equal to the moment of all external
forces about o.
Alternatively, we could formulate the law of angular momentum about
the system’s C.G.. First, differentiate (2.9) with time:
d d d
Ho = Hc + (rc × M vc ) (2.22)
dt dt dt
From (2.22), the first term in (2.22) is rewritten as

d
Ho = Mo
dt 
= r × Fi (2.23)
 i
= (r + ρi ) × Fi
 c 
= rc × Fi + ρi × Fi
40 CHAPTER 2. LINEAR AND ANGULAR MOMENTUMS

Z
ω

e3 e2 ρ dm
A

e1
rA r

o Y

Figure 2.2: A rigid body

In addition, the third term in (2.22) is then

d d d
(rc × M vc ) = rc × M vc + rc × (M vc )
dt dt  dt (2.24)
= vc × M vc + rc × Fi

= 0 + rc × Fi

Substitution of (2.23) and (2.24) into (2.22) yields


 d
ρi × Fi = Hc (2.25)
dt
Or

Ḣc = ρi × Fi = Mc (2.26)

From equation (2.26), if proper coordinates are used to described ρi then


we can define a set of geometric quantities so called moments of inertia of
a rigid body. The moments of inertia measure the angular momentum per
unit rate of rotation. They will be derived in detail in Section 2.3.

2.2 Angular Momentum of a Rigid Body


Figure 2.2 shows a rigid body moving in 3D. Let the angular velocity of the
body be ω. Consider a rigid body as a continuous media of particles with
2.2. ANGULAR MOMENTUM OF A RIGID BODY 41

no deformation, the angular momentum of this rigid body is the integral


form of (2.7) or 
Ho = r × vdm (2.27)

where r = rA +ρ is a position vector from the fixed origin o to the differential


mass dm of the body. Let e1 e2 e3 in Figure 2.2 be the rotating reference
frame with its origin located at an arbitrary point A on the body. If the
reference coordinate system and the body have the same angular velocity,
i.e. ω, then
v = ṙ = vA + ω × ρ (2.28)
Note that the velocity of the differential mass dm relative to e1 e2 e3 is zero
because: 1) the rigid body has no deformation and 2) e1 e2 e3 rotates syn-
chronously with the body. Substitution of (2.28) into (2.27) yields

Ho = [(rA + ρ) × (vA + ω × ρ)] dm
  
= rA × vA dm + ρdm × vA
   
+rA × ω × ρdm + ρ × (ω × ρ) dm

= rA × mvA + mρc × vA + rA × [ω × mρc ] + ρ × (ω × ρ) dm


(2.29)
where m = dm is the mass of the rigid body and ρc is the position of
1 
the body’s C.G. measured with respect to A and given by ρc = m ρdm.
Furthermore, the rotation of a rigid body can be considered as two different
cases: pure rotation and general motion (combined rotation and transla-
tion).

1. Pure rotation about fixed point o


In this case, if we choose point A in Figure 2.2 fixed at o. Hence
rA = vA = 0, and (2.29) becomes

Ho = ρ × (ω × ρ) dm (2.30)

2. General motion
In this case if point A in Figure 2.2 is fixed at the body’s C.G, i.e.
point c. Hence rA = rc and ρc = 0. (2.29) then becomes

Ho = rc × mvc + ρ × (ω × ρ) dm (2.31)
42 CHAPTER 2. LINEAR AND ANGULAR MOMENTUMS

For a rigid body, the angular momentum about c is defined as



Hc ≡ ρ × (ω × ρ) dm (2.32)

Hence
Ho = rc × mvc + Hc (2.33)

2.3 Mass Moment of Inertia


Due to a constant geometric property of the rigid body, the angular mo-
mentum can be more simplified as follows. First it is noted that the angular
momentum about o (2.30), in case of pure rotation, and the angular momen-
tum about c (2.32), in case of general motion, have the same form. Therefore
we will drop out the subscripts in (2.30) and (2.32) for convenience and gen-
erally rewrite both equations as

H= ρ × (ω × ρ) dm (2.34)

Since a triple cross product can be rewritten as A × (B × C) = (A · C)B −


(A · B)C, then (2.34) becomes

H= (ρ · ρ)ω − (ρ · ω)ρdm (2.35)

ρ and ω can be expressed in terms of components with respect to e1 e2 e3


coordinate system as follows
 T  T
ρ= x y z , ω= ωx ωy ωz

Next we will rewrite (2.35) in terms of its components. Let’s consider (2.35)
term by term. The first term is
 
(ρ · ρ)ωdm = (ρ · ρ) [δ] ωdm
 
ρ2 0 0
 
=  0 ρ2 0  ωdm
0 0 ρ2  (2.36)
  
 ρ2 0 0
   
=   0 ρ2 0  dm ω
0 0 ρ 2
2.3. MASS MOMENT OF INERTIA 43

where ρ2 = x2 + y 2 + z 2 and [δ] is the identity matrix. The second term of


(2.35) is

    
   ωx x
    
− (ρ · ω)ρdm = −  x y z  ωy   y  dm
ωz  z
 x
 
= − (xωx + yωy + zωz )  y  dm

z 
 2
x ωx + xyωy + xzωz
 
= −  xyωx + y 2 ωy + yzωz  dm (2.37)

xyωx + yzωy+z 2 ωz 
 x2 xy xz ωx
  
= −  xy y 2 yz   ωy  dm

xz yz z 2  ωz
 x2 xy xz
 
= −  xy y 2 yz  dm(ω)
xz yz z 2

Substitution of (2.36) and (2.37) into (2.35) yields

   
 y2 + z2 −xy −xz
   
H =   −xy 2
x +z 2 −yz  dm ω

−xz 
−yz x2 + y 2
(2.38)
I11 I12 I13
 
=  I21 I22 I23  ω = Iω
I31 I32 I33

where I is the matrix of (second) moments of inertia. The components of I


along diagonal are called moments of inertia given by

  
I11 = y 2 + z 2 dm
  
I22 = x2 + z 2 dm (2.39)
  
I33 = x2 + y 2 dm
44 CHAPTER 2. LINEAR AND ANGULAR MOMENTUMS

and the off-diagonal components so called cross product of inertia are given
by 
I12 = I21 = − xydm

I13 = I31 = − xzdm (2.40)

I23 = I32 = − yzdm

Any three orthogonal axes e1 e2 e3 that yields all zero cross product of inertia,
i.e. I12 = I13 = I23 = 0, are called principal axes. In this case I11 = I1 ,
I22 = I2 , and I33 = I3 are called the principal inertias. I1 , I2 , and I3 can
be determined from the eigenvalues of the matrix I. With the principal
inertias, the angular momentum of a rigid body can be simplified as

H = I1 ω1 e1 + I2 ω2 e2 + I3 ω3 e3 (2.41)

Properties of I

1. I is a symmetric matrix

2. I has positive eigenvalues which are principal inertias I1 , I2 , and I3 ,


and has three orthogonal eigenvectors which represent the principal
axes e1 e2 e3 .

3. For a basis with at least two symmetry planes, the off-diagonal terms
or the cross-product of inertia are zero.

4. The parallel axes theorem states that


(c)
Ikk = Ikk + m∆2k , k = 1, 2, 3 (2.42)

and
(c)
Iij = Iij − mdi dj , i, j = 1, 2, 3, i = j (2.43)

where ∆k is the distance between the two parallel axes, and di and dj
are the relative displacements along i and j coordinates, respectively.

5. The inertia matrix calculation is an additive operator.

Practical methods used to determine the inertia matrix are: 1) look-up table,
2) computer calculation, and 3) experiment.
Chapter 3

Dynamics of a Rigid Body

In this chapter, the dynamics of a rigid body for two different cases: 1) a
pure rotation; and 2) a general motion consisting of both translation and
rotation, are analyzed using the Newton-Euler approach. According to the
laws of angular momentum stated in the previous chapter, we can formulate
the dynamics equations governing the motion of a rigid body.

3.1 Newton-Euler Equations of a rigid body


For a rigid body having pure rotation about o with the angular velocity ω,
the governing equation is 
Mo = Ḣo (3.1)
Equation (3.1) is the law of angular momentum for a rigid body. In this
case, we choose the reference coordinate system {e1 e2 e3 }, with its origin
fixed at o, that rotates with the body with the same angular velocity ω.
Hence, the angular momentum about o can be simplified as

Ho = I o ω (3.2)

where Io is the constant matrix of moments of inertia about o whose com-


ponents are along {e1 e2 e3 } axes.
For a general motion of a rigid body, the equations governing both trans-
lation and rotation are 
F = mv̇c (3.3)
and 
Mc = Ḣc (3.4)
where point c is the C.G. of the rigid body.

45
46 CHAPTER 3. DYNAMICS OF A RIGID BODY

Equation (3.3) is the law of linear momentum for a rigid body or so called
the Newton’s equation, and equation (3.4) is the law of angular momentum.
For the general motion, we normally choose the reference coordinate system
such that its origin is fixed at the C.G. of the body and its coordinates rotate
with the body. If the rigid body has the angular velocity ω, then

Hc = I c ω (3.5)

where Ic is the constant matrix of moments of inertia about c whose com-


ponents are along the reference coordinates.
For both cases of motion, if the reference coordinates, i.e. e1 e2 e3 , are
in the directions such that they are the principal axes, then Ho and Hc in
(3.2) and (3.5) are simply

Ho = Io ω = I1o ω1 e1 + I2o ω2 e2 + I3o ω3 e3 (3.6)

and
Hc = Ic ω = I1c ω1 e1 + I2c ω2 e2 + I3c ω3 e3 (3.7)
Hence (3.1) and (3.4) can be more simplified as
      
 M1o I1o ω̇1 0 −ω3 ω2 I1o ω1
      
Mo =  M2o  =  I2o ω̇2  +  ω3 0 −ω1   I2o ω2 
M3o I3o ω̇3 −ω2 ω1 0 I3o ω3
(3.8)
and
      
 M1c I1c ω̇1 0 −ω3 ω2 I1c ω1
      
Mc =  M2c  =  I2c ω̇2  +  ω3 0 −ω1   I2c ω2 
M3c I3c ω̇3 −ω2 ω1 0 I3c ω3
(3.9)
Or they can be written in a scalar form as

M1o = I1o ω̇1 + (I3o − I2o ) ω2 ω3


M2o = I2o ω̇2 + (I1o − I3o ) ω1 ω3 (3.10)
M3o = I3o ω̇3 + (I2o − I1o ) ω1 ω2

and
M1c = I1c ω̇1 + (I3c − I2c ) ω2 ω3
M2c = I2c ω̇2 + (I1c − I3c ) ω1 ω3 (3.11)
M3c = I3c ω̇3 + (I2c − I1c ) ω1 ω2
Equation sets (3.10) and (3.11) are called Euler’s equations.
3.1. NEWTON-EULER EQUATIONS OF A RIGID BODY 47

u(t)
massless cart
ey

ex m, L
g θ

Figure 3.1: A cart-pendulum system

Example 3.1 : Dynamics of a pendulum-cart system


The cart with a negligible weight moves along a frictionless floor as shown in
Figure 3.1. The pendulum with mass m and length L is hinged to the cart
at one end. If the cart motion is prescribed by u(t), derive the equation of
motion of the system.
Solution:
First, consider the pendulum or the uniform rod which has a general
plane (2-D) motion. The degree of freedom used to describe the motion of
this rod is θ(t).
Kinematics: With the coordinate systems shown in Figure 3.2, the velocity
vc and acceleration ac at C.G. of the rod are
L
vc = u̇(t)ex + θ̇ eθ (3.12)
2
L L
ac = v̇c = ü(t)ex + θ̈ eθ − θ̇ 2 er (3.13)
2 2
With the free body diagram (FBD) shown in Figure 3.2, we set Newton-
Euler equations as

[ F = mv̇c ];
 
L L
Fr er + Fθ eθ − mgey = m ü(t)ex + θ̈ eθ − θ̇ 2 er (3.14)
2 2

and [ Mc = Ic ω̇];
L
Fθ = Ic θ̈ (3.15)
2
48 CHAPTER 3. DYNAMICS OF A RIGID BODY

Fr

ey
eθ eθ
θ c
ex θ
θ
er er
mg

Figure 3.2: Coordinate systems and FBD

Note that Ic in (3.15) is the moment of inertia about the C.G. of the rod
along z-axis. We now have three unknowns: Fr , Fθ , and θ, and three scalar
equations, two from (3.14) and one from (3.15). To derive the equation of
motion, we need to eliminate all unknown forces which are Fr and Fθ and
reduce the Newton-Euler equations to only one differential equation.
Figure 3.2 shows the two coordinate systems with the coordinate trans-
formation given by
ex = sinθer + cosθeθ
(3.16)
ey = −cosθer + sinθeθ
To eliminate Fθ , we substitute (3.15) into (3.14) and transform all coordi-
nates to {er eθ } using (3.16). Then (3.14) can be expressed in scalar form
as:
r-component:
L
mθ̇ 2 + mgcosθ = Fr + müsinθ (3.17)
2
θ-component:
 
mL2 mgL mL
Ic + θ̈ + sinθ = ücosθ (3.18)
4 2 2

Equation (3.18) is the equation of motion. Note that (3.18) is a nonlinear


equation. The solution of (3.18) can be obtained from a numerical integra-
tion using Matlab. Otherwise if only small oscillation is interested, we can
3.2. MODIFIED EULER’S EQUATIONS 49

e2
ρ C.G. e
dm 1
Z e3 Ω
ω0
Y
X

Figure 3.3: A rigid body with symmetric shape

linearize (3.18) to get a closed-form solution. With the solution of (3.18),


the dynamic forces Fr and Fθ are obtained from (3.15) and (3.17) as

L
Fr = mθ̇ 2 + mgcosθ − müsinθ (3.19)
2
and
2Ic
Fθ = − θ̈ (3.20)
L

3.2 Modified Euler’s equations


A set of modified Euler’s equations is used in the case of the symmetric-
shape rigid body which spins about its symmetry axis with a constant speed,
as shown in Figure 3.3. To formulate the Modified Euler’s equations, two
conditions are defined.

1. The rigid body spins about the symmetry axis with a constant speed
ωo .

2. The reference coordinate system e1 e2 e3 is chosen such that one of the


axes, i.e. e3 , is the symmetry axis. In addition, e1 e2 e3 only precesses
but does not spin with the body. Also the origin of e1 e2 e3 is fixed at
the point of rotation for the case of pure rotation, and is fixed at the
50 CHAPTER 3. DYNAMICS OF A RIGID BODY

body’s C.G. for the case of general motion. In these cases, e1 e2 e3 are
principal axes and I1 = I2 ≡ I. If the angular velocity of e1 e2 e3 is

Ω = Ω 1 e1 + Ω 2 e2 + Ω 3 e3

The angular velocity of the rigid body is then

ω b = Ω + ω o e3

According to the conditions above, the modified Euler’s equations become

M1o = Io Ω̇1 + (I3o − Io ) Ω2 Ω3 + I3o ωo Ω2


M2o = Io Ω̇2 + (Io − I3o ) Ω3 Ω1 − I3o ωo Ω1 (3.21)
M3o = I3o Ω̇3

and
M1c = Ic Ω̇1 + (I3c − Ic ) Ω2 Ω3 + I3c ωo Ω2
M2c = Ic Ω̇2 + (Ic − I3c ) Ω3 Ω1 − I3c ωo Ω1 (3.22)
M3c = I3c Ω̇3
Equation sets (3.21) and (3.22) are for the cases of pure rotation and general
motion, respectively. The derivation of the modified Euler’s equations is
shown for the case of general motion as follows. From Figure 3.3, the angular
momentum of the rigid body about its C.G. is

Hc = (ρ × ρ̇)
dm
 dρ

= ρ × dt + Ω × ρ dm
 rel
= 
(ρ × [(ωo e3 × ρ) + Ω × ρ]) dm
= 
(ρ × (ωo e3 + Ω) × ρ) dm (3.23)
= (ρ × 
ω b × ρ) dm   
Ic 0 0 Ω1
  
= Iω b =  0 Ic 0   Ω2 
0 0 I3c Ω3 + ωo

Then
 
Ω̇1
 
Ḣc = I  Ω̇2  + Ω × Hc
 Ω̇3     (3.24)
Ic Ω̇1 0 −Ω3 Ω2 Ic Ω1
    
=  Ic Ω̇2  +  Ω3 0 −Ω1   Ic Ω2 
I3c Ω̇3 −Ω 2 Ω 1 0 I3c (Ω 3 + ω o )
3.2. MODIFIED EULER’S EQUATIONS 51

.
ψ
e3
.
φ
L
k
e1
θ
mg
o

Figure 3.4: A gyro top

Substitution of (3.24) into (3.4) hence results in (3.22).


Example 3.2 : Steady precession of a gyro top
Derive the dynamic equation governing steady precession of a gyro top
shown in Figure 3.4. With the steady precession, the top has constant
precession rate ψ̇ and constant spin rate φ̇, and the nutation angle θ is also
constant.
Method1: Direct approach In Figure 3.4, the angular velocity of the
reference coordinate system e1 e2 e3 is
ω e1 e2 e3 = ψ̇k + θ̇e2 (3.25)
Also the angular velocity of the body is
ωb ≡ ω 1 e1 + ω 2 e2 + ω 3 e3
= ψ̇k + θ̇e2 + φ̇e3
= ψ̇ (cosθe3 + sinθe1 ) + θ̇e2 + φ̇e3 (3.26)
= ψ̇sinθe1 + θ̇e2 + φ̇ + ψ̇cosθ e3

Hence ω1 = ψ̇sinθ, ω2 = θ̇, and ω3 = φ̇ + ψ̇cosθ. Since e3 is the


symmetric axis, therefore I1 = I2 and all cross products of inertia are
zero. As the previous proof in (3.23), it can be similarly shown that
the angular momentum of the body about o is Ho = Io ω b . Or
Ho = I1 ω1 e1 + I1 ω2 e2 + I3 ω3 e3  (3.27)
= I1 ψ̇sinθe1 + I1 θ̇e2 + I3 φ̇ + ψ̇cosθ e3
52 CHAPTER 3. DYNAMICS OF A RIGID BODY

For a steady motion, θ is constant or θ̇ = 0, and ω̇1 = ω̇2 = ω̇3 = 0.


The angular momentum is then
 
Ho = I1 ψ̇sinθe1 + I3 φ̇ + ψ̇cosθ e3 (3.28)

Note that the angular momentum of the steady gyro in (3.28) has a
constant magnitude, and the direction of the angular momentum is on
the plane of rotations, i.e. k − e3 plane. Moreover, the rate of change
of angular momentum is
Ḣo = ψ̇k × Ho (3.29)
Substituting (3.28) into (3.29) and performing a matrix operation yield
 
Ḣo = (I1 − I3 ) ψ̇ 2 sinθcosθ − I3 ψ̇ φ̇sinθ e2 (3.30)
 
From the law of angular momentum Mo = Ḣo , the moment sum

Mo about o, in this case, is due to only the gravitation force, given
by 
Mo = Le3 × −mgk = −mgLsinθe2 (3.31)
Note that moment of the resultant force about o shown in (3.31) is
always perpendicular to both rotation axes (k and e3 ), resulting in
the precession. For the steady precession, this moment has a constant
magnitude due to the constant nutation angle θ. Substitution of (3.30)
and (3.31) into the law of angular momentum yields
−mgLsinθ = (I1 − I3 ) ψ̇ 2 sinθcosθ − I3 ψ̇ φ̇sinθ (3.32)
For sinθ = 0, we get the equation governing a steady precession of the
top as
mgL = I3 ψ̇ φ̇ + (I3 − I1 ) ψ̇ 2 cosθ (3.33)
From (3.33), with a given θ we can determine the relation between the
π
precession rate ψ̇ and the spin rate φ̇. For example, if θ = equation
2
(3.33) becomes
mgL
ψ̇ φ̇ = (3.34)
I3
Method 2: modified Euler’s equations The modified Euler’s equations
are presented again as follows
M1o = Io Ω̇1 + (I3o − Io ) Ω2 Ω3 + I3o ωo Ω2
M2o = Io Ω̇2 + (Io − I3o ) Ω3 Ω1 − I3o ωo Ω1 (3.35)
M3o = I3o Ω̇3
3.3. INTRODUCTION TO STABILITY OF A SPIN BODY 53

y z

Figure 3.5: Stability of a spin plate

In (3.35), Ω1 = ψ̇sinθ, Ω2 = 0, and Ω3 = ψ̇cosθ. Also Ω̇1 = Ω̇2 =


0 because of a steady motion. Furthermore, the spin rate ωo = φ̇.
Substitution of these terms into (3.35) yields

M2o = −mgLsinθ = (I1 − I3 ) ψ̇ 2 sinθcosθ − I3 ψ̇ φ̇sinθ (3.36)

Equation (3.36) is equivalent to (3.32) from Method 1.

3.3 Introduction to stability of a spin body


Stability analysis of a spin plate:
Let xyz be principal axes of a spinning rectangular plate as shown in
Figure 3.5. We want to analyze the stability of rotation about each principal
axis.
By stability of rotation, we ask the question: during a steady spin about
each axis, if the initial rotation is applied so close to the principal axes (is
perturbed a bit in every directions), will the rotation remain close to the
principal axes (does the perturbation die out), or will the body begin to see
increasing rotation about one of the other axes (does the perturbation grow
with time)?
To analyze this problem, let’s first formulate the Euler’s equations for
54 CHAPTER 3. DYNAMICS OF A RIGID BODY

the spin plate as follows:

M1c = I1c ω̇1 + (I3c − I2c ) ω2 ω3


M2c = I2c ω̇2 + (I1c − I3c ) ω1 ω3 (3.37)
M3c = I3c ω̇3 + (I2c − I1c ) ω1 ω2

where subscripts 1, 2, and 3 in (3.37) denote the principal axes of the plate.
Due to a steady spin, the system is moment-free. Hence in (3.37) M1c =
M2c = M3c = 0. Let’s assume that the plate has a steady spin about the
axis ‘1’ with a constant speed ω0 . (Note that axis-1 can be any arbitrary
principal axis, i.e. x-, y-, or z-axis in Figure 3.5.) Then the plate is perturbed
with small angular velocities η1 (t), η2 (t), and η3 (t), respectively, about all
principal axes. Hence the angular velocities in each direction are

ω1 (t) = ω0 + η1 (t)
ω2 (t) = η2 (t) (3.38)
ω3 (t) = η3 (t)

Substitution (3.38) into (3.37) and neglecting the higher order terms, such
as η1 η2 , η2 η3 , etc., yield

I1c η̇1 = 0
I2c η̇2 + (I1c − I3c ) ω0 η3 = 0 (3.39)
I3c η̇3 + (I2c − I1c ) ω0 η2 = 0

The first row of (3.39) implies that η1 (t) is constant. In addition, the last
two rows of (3.39) can be written in a matrix form as
   (I1c −I3c )ω0
   
η̇2 (t) 0 I2c η2 (t) 0
+ (I2c −I1c )ω0 = (3.40)
η̇3 (t) 0 η3 (t) 0
I3c

or
η̇(t) + Kη(t) = 0 (3.41)
To solve (3.40), assume the solution as the following form
   
η2 (t) a
η(t) = = eλt (3.42)
η3 (t) b

Substitution (3.42) into (3.41) yields


   
a λt 0
[λI + K] e = (3.43)
b 0
3.3. INTRODUCTION TO STABILITY OF A SPIN BODY 55

For a nontrivial solution, we get the characteristic equation: |λI + K| = 0.


The characteristic roots λ can be solved as
(I1c − I3c ) (I2c − I1c ) ω02
λ2 = (3.44)
I2c I3c
There are two roots of λ which are
 1
(I1c − I3c ) (I2c − I1c ) ω02 2
λ1,2 =± (3.45)
I2c I3c

With two roots, the solution (3.42) is then


     
η2 a1 λ1 t a2
= e + eλ2 t (3.46)
η3 b1 b2

To analyze the stability from the values of λ, we can divide λ2 into two cases
as follows

Case I: (λ2 ≤ 0) In this case, λ1,2 are positive and negative imaginary parts
and the rotation are marginally stable. Specifically, the perturbation
causes the oscillatory motion about the steady state. To satisfy this
stable condition, I1c > I2c > I3c or I1c < I2c < I3c . In other words, the
moment of inertia about the spin axis I1c should be either maximum
or minimum.

Case II: (λ2 > 0) In this case, one of the root is positive real and the other
is negative real. With the positive real root, the solution (3.46) shows
that the rotation is about to increase exponentially with time and
hence the rotation of the plate is unstable.

From this analysis together with a real demonstration, the students should
be able to figure out that in which directions the rotation of the spin plate
are stable.
56 CHAPTER 3. DYNAMICS OF A RIGID BODY
Chapter 4

Multi-Body Mechanical
System

4.1 Degrees of Freedom (DOF)


Degrees of freedom are a complete set of independent coordinates that used
to describe the motion. For example, a rigid body performing free motion
(without any constraints) in 3-D space needs six degrees of freedom (coor-
dinates) to describe its motion, i.e. three for translations and another three
for rotations. For a system of N -rigid bodies having the 3-D free motion,
the number of DOFs is 6 × N .

4.2 Constraints
If any two rigid bodies are connected to each other, the mechanism con-
necting the bodies is called constraint. The constraint imposes additional
relative motion of one body with respect to anothers. With constraints, the
motion of each rigid body in all six coordinates are not independent, hence
the number of DOF for each body is reduced to less than six.

4.3 Constraint Equations


The constraint equations describe the relative motions of any two connected
bodies. We can learn to construct these constraint equations by the following
examples.

57
58 CHAPTER 4. MULTI-BODY MECHANICAL SYSTEM

φ
z

θz

θx y
A
ry

Rz
Mz

Rx

Mz

Figure 4.1: A slider

Example 1: slider Four constraint forces and couples Rx , Rz , Mx and


Mz in the frictionless slider A as shown in Figure 4.1 result in four
constraint equations, i.e. rx = 0, rz = 0, θx = 0 and θz = 0. Without
friction, the slider translates free along y-direction and also rotate free
about y-axis. In this case, two coordinates such as ry and φ as seen
in Figure 4.1 can be chosen as the DOFs to describe such translation
and rotation.

Example 2: spherical joint Three constraint forces Rx , Ry , and Rz in


the spherical joint as shown in Fig. 4.2 result in three constraint equa-
tions, i.e. rx = 0, ry = 0, and rz = 0. In Figure 4.2, link B that
connected to the stationary link A through the joint can rotate free
about its center, assuming no friction. In this case, three spherical
coordinates or the conventional Euler angles θ, ψ, and φ are the DOFs
used to describe the rotation.

Example 3: rolling sphere Consider the spherical ball rolls without slip-
ping as shown in Fig. 4.3. The first geometric constraint relation, i.e.
rcz = a, can be simply observed. Another two relations are derived
from the fact that the contact point A on the sphere is motionless with
respect to the contact point A on the surface. Hence

(vAx )rel = ṙcx − θ̇y a = 0


4.3. CONSTRAINT EQUATIONS 59

Rz
z
φ
B
ψ
A θ y
Ry
x Rx

Figure 4.2: Ball and socket

z
a
θy
c
θx
y
rcz
x rcx
A
rcy A’

Figure 4.3: A rolling sphere


60 CHAPTER 4. MULTI-BODY MECHANICAL SYSTEM

(vAy )rel = ṙcy + θ̇x a = 0

Or the velocities of the C.G. are then

vcx = ṙcx = θ̇y a

vcy = ṙcy = −θ̇x a

Note that there exist three unknown constraint forces Rx , Ry , and Rz


for this case.

From these previous examples, the number of DOFs of each body is equal
to [6 − number of constraint equations (or constraint forces)]. The chosen
DOFs in each case are called generalized coordinates.
Now let’s consider the multi-body linkages in Fig. 4.4. From the previous
examples, we can conclude that the total constraint equations is equal to 4
(from the slider) + 3 (from the spherical joint) = 7. The number of DOFs
is therefore equal to 2 × 6 − 7 = 5. The generalized coordinates, in this case,
are ry , φ, and the other three spherical coordinates at the spherical joint.
Generally speaking, the number of degrees of freedom of a multi-body
system is

M =6×N − C

where M is the number of degrees of freedom, N is number of rigid bodies,



C is number of all constraint equations.

4.4 Classification of Constraints


If the constraint equation can be derived as a function of only generalized
coordinates and time, e.g. examples 1 and 2 in section 4.3, these constraints
are classified as holonomic constraints. In addition, the holonomic con-
straints can be divided into two classes: scleronomic and rheonomic. The
constraint equation for the scleronomic constraint is an implicit function of
time whereas the equation for rheonomic constraint is an explicit function
of time.
If one of the constraint equation is a function of both the generalized
coordinates and their time derivatives, such constraint is classified as non-
holonomic constraint, e.g. example 3 in Section 4.3: rolling sphere.
4.5. NUMBER OF DOF VS. DRIVING FORCES 61

z
θz
Z
θx
θy
x φ
y

Y
rY

Figure 4.4: Combined constraints

4.5 Number of DOF vs. Driving Forces


If the motion along L coordinates can be prescribed as functions of time, so
called the prescribed motions, the number of DOF is then reduced by L. In
order to have the mechanical system perform such prescribed motions, the
corresponding driving forces need to be applied to the system. For instant,
the driving torque is applied to the motor to assure a constant speed of the
rotor. With the prescribed motion in the system, the number of DOF of
multi-body system is 
M =6×N − C −L
where L is number of the prescribed motions.

4.6 Dynamic Analysis of Multi-Body Mechanical


Systems
Dynamic analysis of a multi-body mechanical system can be separated into
two main parts: kinematics and kinetics. Detailed analysis of each part is
described as follows.

Kinematic analysis :

1. Choose reference coordinate system for each body


62 CHAPTER 4. MULTI-BODY MECHANICAL SYSTEM

2. Define generalized coordinates


3. Formulate components of velocity and angular velocity in terms of
the generalized coordinates along the reference coordinate system

Kinetic analysis :

1. Express Newton-Euler’s equations governing dynamics of each


rigid body
2. With free body diagram (FBD), determine components of forces
and moments corresponding to the reference coordinates
3. Substitute forces and kinematic relations into Newton-Euler’s
equations
4. Eliminate all unknown forces to obtain equations of motion (num-
ber of equations of motion is equal to number of DOF.)
5. Solve the equations of motion to determine the time responses
and then use them to obtain all unknown forces

4.7 Example Problem: Dynamics of Two-Link Arms


The two-link arms are connected by the hinge support A as shown in Fig-
ure 4.5. Link 1 is approximately massless and is driven by a motor which
is excluded from the system. The driving torque Md provided by the motor
is related to the speed ω (in rad/s) as Md = M0 − ∆M ω, where M0 and
∆M are constant parameters. Link 2 has mass m and length l. In addition,
the rest dimensions and coordinates are shown in Fig. 4.5. Derive equation
governing the motion of link 2 and solve for time response, given the initial
conditions: β(0) = 0, β̇(0) = 0, ω(0) = 0, and ω̇(0) = 0.
Kinematic analysis
Number of DOF = (2 × 6) - number of constraint equations = 2 × 6 − (5 + 5)
=2
Therefore we need two DOFs to describe the motion of this system. In
this case we choose α and β as the generalized coordinates. Fig. 4.5 also
shows the coordinate systems and their unit vectors.
The angular velocities of link 1 and 2 and the velocity at the C.G. (point
c) of link 2 are, respectively,

Ω1 = α̇k1 (4.1)
4.7. EXAMPLE PROBLEM: DYNAMICS OF TWO-LINK ARMS 63

z2

y1 MAz
α MAy
Y A
α
y2
X RAy RAx
x1 RAz
c
Z, z1
FBD of link 2
mg

z2
Z, z1
A RAz
β MAy
y1
m, l y2 RAx A
rG

y;
;y
c RAy
Link 1 β MAz
Md
Link 2
Md
C

RCY MCY
a
RCX
MCX RCZ
FBD of link 1

Figure 4.5: Two-link arms


64 CHAPTER 4. MULTI-BODY MECHANICAL SYSTEM

Ω2 = α̇k1 + β̇i2
(4.2)
= β̇i2 + α̇sinβj2 + α̇cosβk2
 
l l
vG2 = −aα̇ − α̇sinβ i2 + β̇j2 (4.3)
2 2
where IJK, i1 j1 k1 , and i2 j2 k2 are the unit vectors of XY Z, x1 y1 z1 , and
x2 y2 z2 , respectively. The acceleration at CG of link 2 is then
 
v̇G2 = −aα̈ − 2l α̈sinβ − 2l α̇β̇cosβ i2
 
+ l
β̈ − aα̇2 cosβ − 2l α̇2 sinβcosβ j2 (4.4)
2 
+ 2l β̇ 2 + aα̇2 sinβ + 2l α̇2 sin2 β k2

Kinetic analysis
Figure 4.5 shows the free body diagram of both links. First let’s consider
link 2. The Newton’s equation governing the translation of link 2 is

mv̇G2 = Fx2 i2 + Fy2 j2 + Fz2 k2 (4.5)

where the resultant forces are determined from the free body diagram as

Fx2 = RAx , Fy2 = RAy − mgsinβ, Fz2 = RAz − mgcosβ (4.6)

Euler’s equations governing the rotation of link 2 are

x2 : M1c = I1c ω̇1 + (I3c − I2c ) ω2 ω3


y2 : M2c = I2c ω̇2 + (I1c − I3c ) ω1 ω3 (4.7)
z2 : M3c = I3c ω̇3 + (I2c − I1c ) ω1 ω2

where
ml2
I1c = I2c = , I3c = 0 (4.8)
12
ω1 = β̇, ω2 = α̇sinβ, ω3 = α̇cosβ (4.9)
l l
M1c = −RAy , M2c = RAx + MAy , M3c = MAz (4.10)
2 2
Substitution of (4.4), (4.6), and (4.8)-(4.10) into (4.5) and (4.7) yields six
scalar equations for link 2:
 
l l
m −aα̈ − α̈sinβ − α̇β̇cosβ = RAx (4.11)
2 2
 
l l
m β̈ − aα̇2 cosβ − α̇2 sinβcosβ = RAy − mgsinβ (4.12)
2 2
4.7. EXAMPLE PROBLEM: DYNAMICS OF TWO-LINK ARMS 65
 
l 2 l
m β̇ + aα̇2 sinβ + α̇2 sin2 β = RAz − mgcosβ (4.13)
2 2
l ml2 ml2 2
−RAy = β̈ − α̇ sinβcosβ (4.14)
2 12 12
l ml2  
RAx + MAy = α̈sinβ + 2α̇β̇cosβ (4.15)
2 12
MAz = 0 (4.16)
Now let’s consider link 1. Since link 1 is massless, all components of the
resultant force and resultant couple are then zero. From FBD of link 1 in
Figure 4.5, consider only the Euler’s equation in z1 -direction which is

z1 : Md − MAz cosβ − MAy sinβ + RAx a = 0 (4.17)

Or
Md RAx a
MAy = − MAz cotβ + (4.18)
sinβ sinβ
Plug (4.18) and (4.16) into (4.15) to obtain

l Md RAx a ml2  
RAx + + = α̈sinβ + 2α̇β̇cosβ (4.19)
2 sinβ sinβ 12

Then plug (4.11) into (4.19) to eliminate RAx and rearrange the equation
as get
2
ma2 α̈ + ml3 α̈sin2 β + malα̈sinβ + 5 2
12 ml α̇β̇sinβcosβ (4.20)
2 α̇β̇cosβ − (M0 − ∆M α̇) = 0
+ mal

To eliminate RAy , plug (4.14) into (4.12) and rearrange the equation as

2 2
lβ̈ − lα̇2 sinβcosβ − aα̇2 cosβ + gsinβ = 0 (4.21)
3 3
Note that (4.20) and (4.21) are the set of equations of motion.
To solve the equations of motion numerically, we rewrite (4.20) and (4.21)
in state form. First, let’s define the state variables x1 = α̇, x2 = β, and
x3 = β̇. By substituting the state variables into (4.20) and (4.21), the
equations of motion can be put into in the state form as follows.
   
ẋ1 f1
   
ẋ =  ẋ2  = f (x1 , x2 , x3 ) =  f2  (4.22)
ẋ3 f3
66 CHAPTER 4. MULTI-BODY MECHANICAL SYSTEM

Time reponses with zero initial conditions


2.5

α (solid) and β (dash); rad/s


2

1.5
.
1

0.5

. 0

-0.5
0 1 2 3 4 5 6 7 8 9 10
time (s)

5
β (degree)

0
0 1 2 3 4 5 6 7 8 9 10
time (s)

Figure 4.6: Time responses of the two link arms

where
M0 − ∆M x1 − (mal/2)x1 x3 cosx2 − (5ml2 /12)x1 x3 sinx2 cosx2
f1 =
ma2 + (ml2 /3)sin2 x2 + malsinx2

f2 = x3
3a 2 3g
x cosx2 − sinx2 f3 = x21 sinx2 cosx2 +
2l 1 2l
Then the state equation (4.22) is numerically solved using Matlab, where
the time response plots is shown in Figure 4.6. Note that the Matlab m-file
is described in Fig. 4.7
4.7. EXAMPLE PROBLEM: DYNAMICS OF TWO-LINK ARMS 67

%Simulation of two-link arms


%x(:,1) is omega
%x(:,2) is beta
%x(:,3) is beta_dot
clear all
t=[0 10]; %initial and final time
x0=zeros(3,1); %initial conditions
[t,x]=ode45('link_eqns',t,x0); %solve nonlinear ode of 2-link arms
figure(1), clf
subplot(2,1,1)
plot(t,x(:,1), 'r', t,x(:,3), 'b--')
xlabel('time (s)')
ylabel('omega (solid) and d(beta)/dt (dash); rad/s')
%grid on
title('Time reponses with zero initial conditions')
subplot(2,1,2)
plot(t,x(:,2)*180/pi)
xlabel('time (s)')
ylabel('beta (degree)')
grid on
%===============================================================
function xdot=link_eqns(t,x)
m= 2; % in kg
a= 0.1; % in m
l= 0.5; % in m
delta_M= 0.5; %in (Nm)sec/rad
M_0= 1; % in Nm
xdot=zeros(3,1);
kk=m*(a^2+l^2/3*sin(x(2))^2+a*l*sin(x(2)));

xdot(1)= (-5/12*m*l^2*x(1)*x(3)*sin(x(2))*cos(x(2)) ...


-m*a*l/2*x(1)*x(3)*cos(x(2)) ...
-delta_M*x(1)+M_0)/kk;
xdot(2)=x(3);
xdot(3)=x(1)^2*sin(x(2))*cos(x(2))+3/2/l*a*x(1)^2*cos(x(2)) ...
-3/2*9.81/l*sin(x(2));

Figure 4.7: Matlab m-file


68 CHAPTER 4. MULTI-BODY MECHANICAL SYSTEM
Chapter 5

Principle of Virtual work

5.1 Virtual Displacement and Virtual Work


A virtual displacement δr is defined as an infinitesimal and instantaneous
displacement in an arbitrary direction that does not oppose or violate con-
straints. Fig. 5.1 and Fig. 5.2 show examples of a particle under constrained
motion. In both figures, F(a) is the applied force, F(c) is the constraint force,
and δr is the virtual displacement. In Fig. 5.1 and Fig. 5.2, the constraint
force F(c) can be expressed as

F(c) = Fc n

where n is a unit vector normal to the path of motion. According the


definition of δr, the virtual displacement is always tangent to the path of
motion and orthogonal to n. Hence

δr · n = 0

As a result, the virtual work δW (c) done by a constraint force is zero, i.e.,

δW (c) ≡ F(c) · δr = Fc n · δr = 0 (5.1)

5.2 Holonomic and Nonholonomic Constraints


For the holonomic constraint, the constraint equation can be derived as a
function of the generalized coordinates and time. Therefore the position
vector can be put in following form

r(t) = f (q1 (t), q2 (t), . . . , qM (t)) (5.2)

69
70 CHAPTER 5. PRINCIPLE OF VIRTUAL WORK

F(a) n
Y

δr
Path of motion
r F(c)

Figure 5.1: Virtual displacement of a particle moving along a constrained


path

g X
θ
δr
(c)
F F(a)

Figure 5.2: Virtual displacement of a bead moving in a circular ring


5.3. GENERALIZED COORDINATES AND JACOBIAN 71

or
r(t) = f (q1 (t), q2 (t), . . . , qM (t), t) (5.3)

where qi (t), i = 1, 2, . . . , M are generalized coordinates and M is the number


of DOFs. Equation (5.2) is for the case of scleronomic constraint where
r(t) is an implicit function of time, and (5.3) is for the case of rheonomic
constraint where r(t) is an explicit function of time.
For the nonholonomic constraint, the constraint equations are functions
of both the generalized coordinates and generalized velocities. Hence the
general form of position vector is given by

r(t) = f (q1 (t), q2 (t), . . . , qM (t), q̇1 (t), q̇2 (t), . . . , q̇M (t), t) (5.4)

Note that most of the following contents, we deal with the holonomic con-
straint.

5.3 Generalized Coordinates and Jacobian


Generalized coordinates denoted by qi , where i = 1, 2, . . . , M 1 , are the set
of independent coordinates used to describe the motion of the system. For a
motion under geometric (holonomic) constraints, qi can be any generalized
variable natural to the constraints. In other words, they are not necessarily
the spatial position variables. The following examples can elaborate this
concept.
The bead moving along the 3-D constrained path
From Fig. 5.3, let’s choose q = s(t) as a generalized coordinate of the bead.
The position vector r of the bead in terms of spatial coordinates is

r = [ x y z ]T

where x, y, and z are spatial position variables. From (5.2), the position
vector r also can be written as a function of the generalized coordinate as

r = r(s)

The virtual displacement δr of the bead is then

dr
δr = δs ≡ jδs (5.5)
ds
1
M is the number of degrees of freedom.
72 CHAPTER 5. PRINCIPLE OF VIRTUAL WORK

Z
s(t)
v(t)

path
r

Figure 5.3: Moving particle along the constrained path

where j is defined a Jacobian vector, representing a unit vector that tangent


to the path of motion, i.e.,
dr(x, y, z) dx dy dz
j= = i+ j+ k
ds ds ds ds
Specifically, the virtual displacement δr is related to the generalized coordi-
nate through this Jacobian. In addition (5.5) can be expressed in a matrix
form as  T
δr = dx ds
dy
ds
dz
ds
δs ≡ Jδs (5.6)
where J is a Jacobian matrix. We can see that the generalized coordinate
s(t) is chosen such that it is tangent to the path or natural to the constraint.
Moving cart with the coin on its inclined plane
Let q1 and q2 in Fig. 5.4 be the generalized coordinates of the coin C. The
position vector r of the coin is given by
r = [ x y z ]T = r (q1 , q2 , t) (5.7)

Note that the position vector r in (5.7) is an explicit function of time because
of the prescribed motion u(t) of the cart which is an explicit function of time.
Virtual displacement δr of the coin is
∂r ∂r ∂r
δr = ∂q1 δq1 + ∂q2 δq2 + ∂t δt
2
∂r (5.8)
= δqi
i=1
∂qi
5.3. GENERALIZED COORDINATES AND JACOBIAN 73

q2

Z C
q1
path
u(t)
r

Figure 5.4: Moving coin on the inclined plane of the moving cart

Note that δt in (5.8) is zero because of the instantaneous virtual displace-


ment. Rewrite δr in terms of spatial coordinates xyz, therefore
2 
 
∂x ∂y ∂z
δr = i+ j+ k δqi (5.9)
i=1
∂qi ∂qi ∂qi
Alternatively, the virtual displacement can be put in a matrix form as
δr = Jδq (5.10)
where  T
δq = δq1 δq1
and J is the Jacobian matrix given by
 ∂x ∂x 
∂q1 ∂q2
 ∂y ∂y 
J =  ∂q 1 ∂q2 
∂z ∂z
∂q1 ∂q2 3×2
A system of particles
Fig. 5.5 shows a system of N -particles with K-geometric constraints. Num-
ber of degree-of-freedom of this system is M = 3N −K. Let q1 , q2 , . . . , qM be
all M generalized coordinates. The position vector r to describe the motion
of all particles is
 T
r = rT1 , rT2 , . . . , rTN
= [x1 , y1 , z1 , x2 , y2 , z2 . . . , xN , yN , zN ]T3N ×1
74 CHAPTER 5. PRINCIPLE OF VIRTUAL WORK

mN m4

m3

Z
rN mj

r1
m1
m2

Figure 5.5: A constrained system of particles

For a system with holonomic constraints, the position vector is also a func-
tion of generalized coordinates:

r = r (q1 , q2 , . . . , qM )

The virtual displacement δr is then

∂r ∂r ∂r
δr = δq1 + δq2 + . . . + δqM
∂q1 ∂q2 ∂qM
M
∂r (5.11)
= δqj
j=1
∂qj

or

M
∂ri
δri = δqj ; i = 1, 2, . . . , N (5.12)
j=1
∂qj

Equation (5.11) can be put in a matrix form as

δr = Jδq (5.13)

where
 T
δq = δq1 δq2 . . . δqM (5.14)
M ×1
5.4. PRINCIPLE OF VIRTUAL WORK 75

and J is the Jacobian matrix given by


 
J11 J12 ... J1M
 
 J21 J22 ... J2M 
J=
 .. .. .. .. 
 (5.15)
 . . . . 
JN 1 JN 2 . . . JN M 3N ×M

∂rk
In (5.15), components of the Jacobian matrix Jkl = for k = 1, 2, . . . , N
∂ql
and l = 1, 2, . . . , M .
In summary, Jacobian embodies the information of unit vectors tangent
to the geometric constraints, which is determined by differentiation of the
physical or spatial variables with respect to the generalized coordinates.
Moreover, the virtual work is related to the generalized coordinates through
this Jacobian. Note that the Jacobians derived in the previous examples are
for the holonomic constraints that satisfy (5.2) or (5.3).

5.4 Principle of Virtual Work


Let’s consider a system of N -particles and M -degrees of freedom. If the
system is in equilibrium then

N
Fi = 0 (5.16)
i=1

where Fi is the total forces acting on the i-th particle. Fi can be divided into
(a)
three types of forces: 1) applied or external forces Fi ; 2) internal spring or
damping forces between ith and j th particles fij ; and 3) constraint forces2
(c)
Fi . Therefore
(a) 
N
(c)
Fi = Fi + fij + Fi , i = j (5.17)
j=1
(a)
The applied forces Fi and the internal spring or damping forces fij can
(ac)
be grouped as working forces so called active forces Fi . In cases of no
(c)
friction and plastic deformation at constraints, the constraint forces Fi are
workless forces. According to (5.1), the virtual work done by all constraint
forces is always zero, i.e.

N
(c)
Fi · δri = 0 (5.18)
i=1
2
Constraints in this case excludes the springs and dampers
76 CHAPTER 5. PRINCIPLE OF VIRTUAL WORK

If the system of particles is in equilibrium, the virtual work δW done by all


forces given by


N
δW = Fi · δri = 0
i=1 (5.19)
N  
(ac) (c)
= Fi + Fi · δri = 0, i = j
i=1

The virtual work δW is zero because of the zero sum of all forces. With the
relation (5.18), (5.19) is simply


N
(ac)
δW (ac) = Fi · δri = 0 (5.20)
i=1

Equation (5.20) is the principle of virtual work stating that if the system
is in equilibrium then virtual work done by all active forces in the system is
zero.

5.5 D’Alembert principle


Let’s consider a system of N -particles and M -degrees of freedom. The New-
ton’s second law applied to such system can be rewritten as the following
alternative form

N
(Fi − mi r̈i ) = 0; (5.21)
i=1

where Fi is the total forces acting on the i-th particle, mi is the mass of the
i-th particle, and r̈i is the acceleration of the i-th particle. The term −mi r̈i
represents an inertia force resisting the motion of the system. From (5.21),
virtual work δW done by all forces, including the inertia force −mi r̈i , is
then zero. Or

N
δW = (Fi − mi r̈i ) · δri = 0 (5.22)
i=1

Since the virtual work done by the constraint forces is always zero, (5.22) is
simply
N 
 
δW  =
(ac)
Fi − mi r̈i · δri = 0 (5.23)
i=1
5.5. D’ALEMBERT PRINCIPLE 77

stating that the virtual work δW  done by all active forces and inertia forces
is zero. For the holonomic constraints, we can substitute the Jacobian rela-
tion (5.12) into (5.23) to get

N 
  
M
(ac) ∂ri
Fi − mi r̈i · δqk = 0 (5.24)
i=1 k=1
∂qk

Rewrite(5.24) as
N

M  (ac) ∂ri
(Fi − mi r̈i ) · δqk = 0 (5.25)
k=1 i=1
∂qk

Since each δqk is independent and nonzero, therefore to satisfy (5.25), each
k-term in the bracket must be zero, i.e.
N 
  ∂r
(ac) i
Fi − mi r̈i · = 0; k = 1, 2, . . . , M (5.26)
i=1
∂qk

∂ri
For a short notation, we define β ik ≡ for the rest of the chapter. Equa-
∂qk
tions (5.26) are the D’Alembert principle that automatically yield the equa-
tions of motion. Again the D’Alembert principle can only be applied to
dynamics of the systems with holonomic constraints.
Example: Direct application of D’Alembert principle
The sliders of masses m1 and m2 are constrained by springs and move along
the frictionless disk slot as shown in Figure 5.6. The disk also rotates about
its center with angular displacement θ. If the unstretched length of the
springs is a, derive the equations of motion.
From the system shown in Figure 5.6, let’s choose ρ1 , ρ2 , and θ as the
generalized coordinates. Hence
 T
q= ρ1 ρ2 θ (5.27)

The system has three degrees of freedom or M = 3. The system consists of


two particles or N = 2. The position vectors of the sliders 1 and 2 are

r1 = f (ρ1 , ρ2 , θ) = ρ1 er (5.28)

and
r2 = f (ρ1 , ρ2 , θ) = −ρ2 er (5.29)
78 CHAPTER 5. PRINCIPLE OF VIRTUAL WORK

Y

er
m2
θ k
θ
k X
m1
ρ2
ρ1

Figure 5.6: Example

The accelerations of both sliders are


   
r̈1 (ρ1 , ρ2 , θ) = ρ̈1 − θ̇ 2 ρ1 er + ρ1 θ̈ + 2ρ̇1 θ̇ eθ (5.30)

and    
r̈2 (ρ1 , ρ2 , θ) = −ρ̈2 + θ̇ 2 ρ2 er − ρ2 θ̈ + 2ρ̇2 θ̇ eθ (5.31)
The active spring forces acting on both sliders are
(ac)
F1 = −k (ρ1 − a) er (5.32)

and
(ac)
F2 = k (ρ2 − a) er (5.33)
The components of the Jacobian matrix can be formulated as follows
∂r1
β 11 ≡ = er
∂ρ1
∂r1
β 12 ≡ = 0
∂ρ2
∂r1
β 13 ≡ = ρ1 eθ
∂θ
∂r2
β 21 ≡ = 0
∂ρ1
∂r2
β 22 ≡ = −er
∂ρ2
5.5. D’ALEMBERT PRINCIPLE 79

∂r2
β 23 ≡ = −ρ2 eθ
∂θ
Substitution of (5.30) to (5.33) into (5.26) yields three sets of equations:
For k = 1,
   
(ac) (ac)
F1 − m1 r̈1 · β 11 + F2 − m2 r̈2 · β 21 = 0 (5.34)

or  
m1 ρ̈1 − θ̇ 2 ρ1 + K (ρ1 − a) = 0 (5.35)
For k = 2,
   
(ac) (ac)
F1 − m1 r̈1 · β 12 + F2 − m2 r̈2 · β 22 = 0 (5.36)

or  
m2 ρ̈2 − θ̇ 2 ρ2 + K (ρ2 − a) = 0 (5.37)
For k = 3,
   
(ac) (ac)
F1 − m1 r̈1 · β 13 + F2 − m2 r̈2 · β 23 = 0 (5.38)

or    
m1 ρ1 θ̈ + 2ρ̇1 θ̇ ρ1 + m2 ρ2 θ̈ + 2ρ̇2 θ̇ ρ2 = 0 (5.39)

Equations (5.35) and (5.37) are the equations of motions. In addition, (5.39)
can be arranged as
d  
m1 ρ21 θ̇ + m2 ρ22 θ̇ = 0 (5.40)
dt
which indicates the conservation of angular momentum of the system or
Ḣo = 0.
80 CHAPTER 5. PRINCIPLE OF VIRTUAL WORK
Chapter 6

Lagrange Mechanics

6.1 Kinetic Energy


Kinetic energy of a system of particles is the total sum of kinetic energy of
each particle given by
1 N
T = mi vi · vi (6.1)
2 i=1

where N is number of particles and mi and vi are the mass and the velocity
of the i-th particle. Kinetic energy of a rigid body as shown in Figure 6.1 is
therefore 
1
T = v · vdm (6.2)
2
where v is the velocity of element dm. From Figure 6.1, the velocity v of
mass dm is
v = vc + ω × ρ (6.3)

Substitution of (6.3) into (6.2) yields1


1
T = 2  (vc + ω × ρ) · (vc + ω × ρ) dm 
2 vc · v dm + vc · ω × ρdm + 12 × ρ) · (ω × ρ) dm
1
= c


(6.4)
2 vc · vc dm + (vc × ω) · ρdm + 2 ω · (ρ × ω × ρ) dm
1 1
=
 
The terms ρdm = 0 and dm = m. Also
 
1 1
ω · (ρ × ω × ρ) dm = ω · ρ × (ω × ρ) dm
2 2
1
Note that (A × B) · C = A · (B × C)

81
82 CHAPTER 6. LAGRANGE MECHANICS

Z
ω

vc
C
ρ

rc

Figure 6.1: Kinetic energy of a rigid body


where ρ × (ω × ρ) dm = Hc is the angular momentum about the C.G.
Hence kinetic energy of a rigid body (6.4) is simply
1 1
T = mvc · vc + Hc · ω (6.5)
2 2
Equation (6.5) can be put in the matrix form as
1 1
T = mvcT vc + ω T Ic ω (6.6)
2 2

6.2 Potential Energy


Some active forces such as gravitational forces and internal forces due to
elastic deformation can be represented by a gradient operator of a scalar
potential energy function V as

F = −∇V (r) (6.7)

where ∇ is the gradient operator given (in cartesian coordinates) by

∂ ∂ ∂
∇= i+ j+ k
∂x ∂y ∂z
6.2. POTENTIAL ENERGY 83

Such forces in form of (6.7) are called conservative forces and the function
V is called potential energy. Consequently work done by the conservative
forces is is independent to the path of motion and equal to the change of the
potential energy ∆V . Therefore the work done by the conservative force F,
resulting in any path of motion from A to B, is
 B
W = F · δr ≡ V (rA ) − V (rB ) (6.8)
A

Or virtual work δW done by the conservative force is

δW = F · δr = −dV (6.9)

Examples of the conservative forces F and their corresponding potential


energy V are as follows.

1. Gravity near the Earth’s surface

F = −mgez , V = mgz

2. Gravitational force
GM m GM m
F=− er , V =−
r2 r

3. Elastic spring force


1 2
F = −kx, V = kx
2

4. Magnetic force between two wires

µ0 I1 I2 µ0 I1 I2 logr
F = , V =−
2πr 2π
where µ0 = 4π × 10−7 in mks units.

5. Electric force between two charges

Q1 Q2 Q1 Q2
F = , V =−
4πε0 r 2 4πε0 r
1
where = 8.99 × 109 in mks units.
4πε0
84 CHAPTER 6. LAGRANGE MECHANICS

6.3 Remarks on Properties of Generalized Coor-


dinates for the System with Holonomic con-
straints
1. For a multi-degree-of-freedom dynamical system, all generalized coor-
dinates q1 , q2 , . . . , qM (where M is the number of degrees of freedom)
are independent.
2. For a system of particles with holonomics (geometric) constraints, the
position vectors ri ; i = 1, 2, . . . , N (N is number of particles) can be
expressed in terms of the generalized coordinates and time t as

ri = ri (q1 , q2 , . . . , qM , t)

3. For a system of particles with holonomics (geometric) constraints, the


virtual displacement can be expressed in terms of the generalized co-
ordinates and time as
∂ri ∂ri ∂ri ∂ri
δri = δq1 + δq2 + . . . + δqM + δt
∂q1 ∂q2 ∂qM ∂t
M
∂ri
= δqj
j=1
∂qj

Note that δt is zero because the virtual displacement is an instanta-


∂ri
neous displacement, and = f (q1 , q2 , . . . , qM , t).
∂qj
4. With only the holonomic constraints in the system, the actual velocity
of the i-th particle can be derived in terms of generalized coordinates
as
∂ri dq1 ∂ri dq2 ∂ri dqM ∂ri
ṙi = + + ... + +
∂q1 dt ∂q2 dt ∂qM dt ∂t
M
∂ri ∂ri
= q̇j +
j=1
∂q j ∂t
≡ f (q1 , q2 , . . . , qM , q̇1 , q̇2 , . . . , q̇M , t)

dqj
where q̇j ≡ is the generalized velocity.
dt
5. Remarkable observation 1
∂ ṙi ∂ri
= , k = 1, 2, . . . , M ; i = 1, 2, . . . , N
∂ q̇k ∂qk
6.4. DERIVATION OF LAGRANGE’S EQUATIONS 85

6. Remarkable observation 2
 
∂ ṙi d ∂ri
= , k = 1, 2, . . . , M ; i = 1, 2, . . . , N
∂qk dt ∂qk

6.4 Derivation of Lagrange’s equations


Let’s consider a system of N -particles with M -degrees of freedom. From the
D’Alembert principle:
N 
  ∂r
(ac) i
mi r̈i − Fi · = 0; k = 1, 2, . . . , M (6.10)
i=1
∂qk

Consider (6.10) term by term. The first term on the left can be written as

N N     
∂ri d ∂ri d ∂ri
mi r̈i · = mi ṙi · − mi ṙi · (6.11)
i=1
∂qk i=1
dt ∂qk dt ∂qk

With the remarkable observation 1 and 2, (6.11) can be rearranged as



N N 
   
∂ri d ∂ ṙi ∂ ṙi
mi r̈i · = mi ṙi · − mi ṙi ·
i=1
∂qk i=1
dt ∂ q̇k ∂qk
N 
     
d ∂ 1 ∂ 1
= mi ṙi · ṙi − mi ṙi · ṙi
i=1
dt ∂ q̇k 2 ∂qk 2
N    
d ∂Ti ∂Ti
= −
i=1
dt ∂ q̇k ∂qk
 
d ∂T ∂T
= −
dt ∂ q̇k ∂qk
where Ti is the kinetic energy of i-particle, and T is the total kinetic energy.
The second term of (6.10) is defined as generalized forces Qk . Or

N
(ac) ∂ri
Qk ≡ Fi · ; k = 1, 2, . . . , M (6.12)
i=1
∂qk

Consequently (6.10) becomes


 
d ∂T ∂T
− = Qk ; k = 1, 2, . . . , M (6.13)
dt ∂ q̇k ∂qk
There are two alternative ways to determine the generalized forces Qk as
described in the following details.
86 CHAPTER 6. LAGRANGE MECHANICS

1. From (6.12), Qk is defined as the component of all active forces that


projected on the direction of k-generalized coordinate.

2. Since virtual work done by all active forces is


N 

M 
(ac) ∂ri M
δW (ac) = Fi · δqk = Qk δqk (6.14)
k=1 i=1
∂qk k=1

Any generalized force Qj is therefore determined from the virtual work


done by enforcing virtual displacement in the particular j-coordinate
for one unit (δqj = 1), and enforcing zero virtual displacements in the
other coordinates (δqi = 0, i = j).

In addition, the generalized forces can be separated into two cases: conser-
(c) (nc)
vative generalized forces Qk and nonconservative generalized forces Qk .
(c)
The conservative generalized force Qk can be expressed in term of the po-
tential energy as

(c) 
N
(c) ∂ri
Qk = Fi ·
i=1
∂qk
N  
∂V ∂ri
= − (q1 , q2 , . . . , qM ) ·
i=1
∂ri ∂qk
∂V
= −
∂qk

(c)
Substitution of the conservative generalized forces Qk into (6.13) yields the
Lagrange’s equations
 
d ∂T ∂T ∂V (nc)
− + = Qk ; k = 1, 2, . . . , M (6.15)
dt ∂ q̇k ∂qk ∂qk

Let’s define a Lagrange function L such that

L(q, q̇) ≡ T (q, q̇) − V (q)

An alternative form of the Lagrange’s equations is then


 
d ∂L ∂L (nc)
− = Qk ; k = 1, 2, . . . , M (6.16)
dt ∂ q̇k ∂qk
6.5. EXAMPLES 87

g l
θ

Figure 6.2: A pendulum

6.5 Examples
Example 6.1 :
Derive the equation of motion of the pendulum in Figure 6.2 using the
Lagrange’s equation.
Solution
The pendulum shown in Fig. 6.2 has one degree of freedom. Let’s choose
θ as the generalized coordinate. The kinetic energy T of the pendulum is
then
1 1
T = mv 2 = ml2 θ̇ 2 (6.17)
2 2
Also the potential energy V of the pendulum (with respect to the datum at
the hinge level) is
V = −mglcosθ (6.18)
There is no external forces applied to the pendulum, hence the generalized
force is zero. The Lagrange’s equation for the pendulum is
 
d ∂T ∂T ∂V
− + =0 (6.19)
dt ∂ θ̇ ∂θ ∂θ

Substitution of (6.17) and (6.18) into (6.19) yields the equation of motion:

ml2 θ̈ + mglsinθ = 0 (6.20)

Example 6.2 :
Derive the equations of motion for the cart-pendulum as shown in Fig. 6.3.
Solution
88 CHAPTER 6. LAGRANGE MECHANICS

x
k
m1


g θ

m2

er

Figure 6.3: A cart and pendulum

The cart-pendulum system has two degrees of freedom. Let the generalized
coordinates be q1 = x, q2 = θ. Absolute velocity v2 of the pendulum is

v2 = ẋex + Lθ̇eθ  (6.21)


= ẋsinθer + ẋcosθ + Lθ̇ eθ

where L is the length of the pendulum. Also


 2
v2 · v2 = ẋ2 sin2 θ + ẋcosθ + Lθ̇
(6.22)
= ẋ2 + 2Lẋθ̇cosθ + L2 θ̇ 2

The total kinetic energy T of the system is

1 1
T = m1 ẋ2 + m2 v2 · v2 (6.23)
2 2
Substitute (6.22) into (6.23) to get

1 1
T = (m1 + m2 ) ẋ2 + m2 Lẋθ̇cosθ + m2 L2 θ̇ 2 (6.24)
2 2
The potential energy of the system is

1 2
V = kx − m2 gL cos θ (6.25)
2
6.5. EXAMPLES 89

The Lagrange’s equations are


 
d ∂T ∂T ∂V
− + =0 (6.26)
dt ∂ ẋ ∂x ∂x
and  
∂T ∂T
d ∂V
− + =0 (6.27)
∂ θ̇ dt
∂θ ∂θ
Formulate each term in (6.26) and (6.27) as follows:
∂T
= (m1 + m2 ) ẋ + m2 Lθ̇cosθ (6.28)
∂ ẋ
 
d ∂T
= (m1 + m2 ) ẍ + m2 Lθ̈cosθ − m2 Lθ̇ 2 sinθ (6.29)
dt ∂ ẋ
∂T
=0 (6.30)
∂x
∂T
= m2 Lẋcosθ + m2 L2 θ̇ (6.31)
∂ θ̇
 
d ∂T
= m2 Lẍcosθ − m2 Lẋθ̇sinθ + m2 L2 θ̈ (6.32)
dt ∂ θ̇
∂T
= −m2 Lẋθ̇sinθ (6.33)
∂θ
∂V
= kx (6.34)
∂x
∂V
= m2 gLsinθ (6.35)
∂θ
Plug (6.29), (6.30) and (6.32) to (6.35) into (6.26) and (6.27) to obtain the
equations of motion:

(m1 + m2 ) ẍ + m2 Lθ̈cosθ − m2 Lθ̇ 2 sinθ + kx = 0 (6.36)

and
m2 L2 θ̈ + m2 Lẍcosθ + m2 gLsinθ = 0 (6.37)
Lagrange’s equations can be used to derive equations of motion of the
rigid body or multi-body system. In this case, derivation of the Lagrange’s
equation is similar to that for the system of particles and will be omitted
here. Also the Lagrange equations for the rigid body or multi-body system
are identical to (6.15) and (6.16).
Example 6.3 :
90 CHAPTER 6. LAGRANGE MECHANICS

g
τ1 a1
m1,
I1(about G) α1
G1
a2
l1
τ2 α2 G2

m2, l2
I2(about G)

Figure 6.4: A rigid two link arm system

The two-link arm robot as shown in Figure 6.4 is operated in a horizontal


plane. The motion of the arms is controlled by two motors installed at the
joints. The motors generate moments τ1 and τ2 as shown in Figure 6.4.
Derive equations of motion for the two-link arm robot.
Solution
Kinematics: There are two degrees of freedom in this case. Let’s define α1
and α2 as the generalized coordinates. The position vector and the velocity
of G2 can be written in terms of both generalized coordinates as
rG2 = (l1 cosα1 + a2 cosα2 ) i + (l1 sinα1 + a2 sinα2 ) j (6.38)
ṙG2 = (−l1 α̇1 sinα1 − a2 α̇2 sinα2 ) i + (l1 α̇1 cosα1 + a2 α̇2 cosα2 ) j (6.39)
The dot product of the velocity ṙG2 is then
ṙG2 · ṙG2 = (−l1 α̇1 sinα1 − a2 α̇2 sinα2 )2 + (l1 α̇1 cosα1 + a2 α̇2 cosα2 )2
= l12 α̇21 + a22 α̇22 + 2l1 a2 α˙1 α˙2 cos (α1 − α2 )
(6.40)
Kinetic and potential energy:
The total kinetic energy T of the two-link arms is
T = T1 + T2  
= 12 Io1 α̇21 + 12 m2 ṙG2 · ṙG2 + 12 IG2 α̇22
  ! "
= 12 I1 + m1 a21 α̇21 + 12 m2 l12 α̇21 + a22 α̇22 + 2l1 a2 α˙1 α˙2 cos (α1 − α2 )
+ 12 I2 α̇22
(6.41)
6.5. EXAMPLES 91

τ1

α1
g
τ2

τ2
α2

Figure 6.5: Reaction torques

Since the whole system operates in the horizontal plane and there is no
restoring forces, the potential energy is zero, i.e. V = 0.
Generalized forces Qα1 and Qα2 :
From Figure 6.5, the virtual work done by all nonconservative torques is

δW = τ1 δα1 − τ2 δα1 + τ2 δα2 (6.42)

Hence
Qα1 = τ1 − τ2 (6.43)
and
Qα2 = τ2 (6.44)
Formulate the Lagrange’s equations as follows:
 
d ∂T ∂T ∂V
− + = Qα1 (6.45)
dt ∂ α̇1 ∂α1 ∂α1
 
d ∂T ∂T ∂V
− + = Qα2 (6.46)
dt ∂ α̇2 ∂α2 ∂α2
where
∂T  
= I1 + m1 a21 α̇1 + m2 l12 α̇1 + m2 a2 l1 α̇2 cos (α1 − α2 ) (6.47)
∂ α̇1
 
d ∂T  
= I1 + m1 a21 + m2 l12 α̈1 + m2 a2 l1 α̈2 cos (α1 − α2 )
dt ∂ α̇1 (6.48)
−m2 a2 l1 α̇2 (α̇1 − α̇2 ) sin (α1 − α2 )
92 CHAPTER 6. LAGRANGE MECHANICS

∂T
= −m2 a2 l1 α˙1 α˙2 sin (α1 − α2 ) (6.49)
∂α1
∂T
= m2 a22 α̇2 + m2 a2 l1 α̇1 cos (α1 − α2 ) + I2 α̇2 (6.50)
∂ α̇2
 
d ∂T  
= m2 a22 + I2 α̈2 + m2 a2 l1 α̈1 cos (α1 − α2 )
dt ∂ α̇2 (6.51)
−m2 a2 l1 α̇1 (α̇1 − α̇2 ) sin (α1 − α2 )
∂T
= m2 a2 l1 α˙1 α˙2 sin (α1 − α2 ) (6.52)
∂α2
Substitution of (6.47)-(6.52) into (6.45) and (6.46) yields the following equa-
tions of motion:
 
I1 + m1 a21 + m2 l12 α̈1 + m2 a2 l1 α̈2 cos (α1 − α2 )
(6.53)
+m2 a2 l1 α̇22 sin (α1 − α2 ) = τ1 − τ2
 
m2 a22 + I2 α̈2 + m2 a2 l1 α̈1 cos (α1 − α2 )
(6.54)
−m2 a2 l1 α̇21 sin (α1 − α2 ) = τ2
Example 6.4 :
Fig. 6.6 shows a uniform and thin bar of mass m and length l hinged to link
1 which is driven to spin with a constant speed ω. Derive the differential
equations governing the motion of the thin bar using Lagrange’s equations.
Solution
With the prescribed motion ω is constant, the number of degrees of freedom

is M = 6 × N − C − L = 6 × 2 − (5 + 5) − 1 = 1. Let’s choose β as the
generalized coordinate. The angular velocity of the link 2 is

ω 2 = ωk1 + β̇i2
= β̇i2 + ωsinβj2 + ωcosβk2 (6.55)
 T
≡ β̇ ωsinβ ωcosβ

Kinetic energy T is
1 1
T = IZ ω 2 + ω T2 Iω2 (6.56)
2 2
Substituting (6.55) into (6.56) yields
  
  I 0 0 β̇
1 2 1   
T = 2 IZ ω + 2 β̇ ωsinβ ωcosβ  0 I 0   ωsinβ 
 
0 0 0 ωcosβ
1 2 1
= 2 IZ ω + 2 β̇ 2 + ω 2 sin2 β
(6.57)
6.5. EXAMPLES 93

k1
ω
k2
j2

g
l

Figure 6.6: A spinning pendulum

Potential energy is V = −mg(l/2)cosβ. The Lagrange’s equation can be


formulated as  
d ∂T ∂T ∂V
− + =Q (6.58)
dt ∂ β̇ ∂β ∂β

where
∂T
= I β̇ (6.59)
∂ β̇
 
d ∂T
= I β̈ (6.60)
dt ∂ β̇

∂T
= Iω 2 sinβcosβ (6.61)
∂β

∂V l
= mg sinβ (6.62)
∂β 2
and Q = 0. Plug (6.59)-(6.62) into (6.58) to get the equation of motion:

l
I β̈ − Iω 2 sinβcosβ + mg sinβ = 0 (6.63)
2
94 CHAPTER 6. LAGRANGE MECHANICS

6.6 Lagrange Multiplier


Lagrange equation is derived from the D’Alembert principle. Originally, it
can be put in the variational form as

M #
   $
d ∂T ∂T ∂V (nc)
− + − Qk δqk = 0; k = 1, 2, . . . , M
k=1
dt ∂ q̇k ∂qk ∂qk
(6.64)
where M is number of degrees of freedom. If all δqk are independent, each
bracket in (6.64) is zero and we obtain the Lagrange equations as shown
in (6.15). If additional p constraints are introduced later, and result in the
dependency of some qk , what happens? Let’s consider the equation (6.64).
The introduction of new constraints will lead to the following conditions:

1. There exist new constraints which can be expressed by the general


form of p constraint equations


M
aik dqk = 0; i = 1, 2, . . . , p (6.65)
k=1

where p is the number of additional constraints. (6.65) is also a general


form of nonholonomic constraints.

2. With conditions (6.65), Equation (6.64) can be rewritten as

M #   $ 
 d ∂T ∂T ∂V (nc) p M
− + − Qk δqk + λi aik δqk =0
k=1
dt ∂ q̇k ∂qk ∂qk i=1 k=1
(6.66)
where λi is called Lagrange Multiplier. Equation (6.66) can be rear-
ranged as
  

M
d ∂T ∂T ∂V (nc)  p
− + − Qk + λi aik δqk = 0 (6.67)
k=1
dt ∂ q̇k ∂qk ∂qk i=1

Now we have M equations from (6.67) plus p constraint equations from


(6.65) and have M + p unknowns which are q1 , q2 , . . . , qM , λ1 , λ2 , . . . , λp . If
the virtual generalized coordinates are arranged such that δq1 , δq2 , . . . , δqM −p
are independent, and δqM −p+1 , δqM −p+2 , . . . , δqM are dependent. Then the
following procedure is performed to derive the equations of motion. First
6.6. LAGRANGE MULTIPLIER 95

we choose λ1 , λ2 , . . . , λp so that each coefficient in the bracket in (6.67) cor-


responding to δqM −p+1 , δqM −p+2 , . . . , δqM is zero. Or
 
(nc) 
p
d ∂T ∂T ∂V
− + −Qk + λi aik = 0; k = M −p+1, M −p+2, . . . , M
dt ∂ q̇k ∂qk ∂qk i=1
(6.68)
With the chosen λi and independent q1 , q2 , . . . , qM −p , the rest coefficients in
the the bracket of (6.67) corresponding to δq1 , δq2 , . . . , δqM −p are all zero.
In summary, we obtain the following relation:
   p
d ∂T ∂T ∂V (nc)
− + = Qk − λi aik ; k = 1, 2, . . . , M (6.69)
dt ∂ q̇k ∂qk ∂qk i=1

Note that the new constraints are introduced to the Lagrange equations as
the generalized forces as seen from the second term on the right of (6.69).
Moreover the Lagrange equation with Lagrange Multiplier can deal with
dynamics with nonholonomic constraints.
Then we solve (6.69) together with the revised constraint relations (6.65),
putting in the form of


M
aik q̇k = 0; i = 1, 2, . . . , p (6.70)
k=1

or in the integral form

fi (q1 , q2 , . . . , qM ) = 0; i = 1, 2, . . . , p (6.71)

Example 6.5 :
Derive equation of motion of a pendulum shown in Figure 6.2 using the
Lagrange multiplier.
Solution
First if we assume that the pendulum is not constrained in the radial direc-
tion, i.e. l is not fixed, this system will have two degrees of freedom. Let
r and θ be the generalized coordinates as shown in Figure 6.7. The kinetic
and potential energies are

1  
T = m ṙ 2 + r 2 θ̇ 2
2

V = −mgr cos θ
96 CHAPTER 6. LAGRANGE MECHANICS

g r
θ

Figure 6.7: A pendulum without constraint

The Lagrange equation in r-coordinate is


 
d
dt
∂T
− ∂T
∂ ṙ
∂V
∂r + ∂r = 0 (6.72)
d
dt (mṙ) − mr θ̇ 2 − mg cos θ = 0

The Lagrange equation in θ-coordinate is


 
d ∂T
− ∂T
+ ∂V
∂θ = 0
dt  ∂ θ̇  ∂θ (6.73)
d
dt mr 2 θ̇ + mgr sin θ = 0

Then we impose the constraint equation, i.e. r = l or δr = 0. Combine


(6.72) and (6.73) together with the imposed constraint, we get
   
d  2 
mr̈ − mr θ̇ − mg cos θ δr
2
mr θ̇ + mgr sin θ δθ + λδr = 0 (6.74)
dt
or
   
d  2 
mr̈ − mr θ̇ 2 − mg cos θ + λ δr mr θ̇ + mgr sin θ δθ = 0 (6.75)
dt

where λ is the Lagrange multiplier. Choose λ such that

mr̈ − mr θ̇ 2 − mg cos θ + λ = 0 (6.76)

and
d  2 
mr θ̇ + mgr sin θ = 0 (6.77)
dt
6.6. LAGRANGE MULTIPLIER 97

In (6.76) and (6.77), there are three unknowns: r, θ and λ. Therefore to


solve these equations for r, θ and λ, we need another one equation which is
the constraint equation:
r=l (6.78)
Plugging (6.78) into (6.76) and (6.77) yields

mlθ̇ 2 + mg cos θ = λ (6.79)

and
ml2 θ̈ + mgl sin θ = 0 (6.80)
Note that (6.80) is equivalent to the equation of motion that we obtain in
Example 6.1. In addition (6.79) gives us the Lagrange multiplier λ which is,
in this case, the tension or the constraint force in the string.
98 CHAPTER 6. LAGRANGE MECHANICS
Chapter 7

Stability Analysis

7.1 Equilibrium, Quasi-Equilibrium, and Steady


States
Equilibrium is the state in which a system is at stationary; i.e. q̇ = 0 and
q̈ = 0, where q is the vector of generalized coordinates.
Quasi-equilibrium or steady state is the state that some coordinates of a
system are in equilibrium, meanwhile the system has steady rotations (with
constant speed) about some axes, e.g. steady precession of the top.

7.2 Stability of Equilibrium or Steady State


To determine if the equilibrium or the steady state is stable, we initially
perturb the system from each state with a small perturbation, and then
investigate how the perturbation changes with time. If the perturbation
dies out or possesses a small oscillation, the perturbed state is stable. If the
perturbation grows with time, that perturbed state is unstable.
From the previous chapters, the equations of motion can be put in a
general form as
q̈i = fi (q1 , q2 , . . . , qk , q̇1 , q̇2 , . . . , q̇k , t) ; i = 1, 2, . . . , k (7.1)
where k is the number of degrees of freedom. Let’s define the state variables
X1 , X2 , . . . , Xk , Xk+1 , Xk+2 , . . . , X2k as
X1 = q1 , X2 = q2 , . . . , Xk = qk (7.2)
and
Xk+1 = q̇1 , Xk+2 = q̇2 , . . . , X2k = q̇k (7.3)

99
100 CHAPTER 7. STABILITY ANALYSIS

Also define a state vector X as

X = [X1 , X2 , . . . , Xk , Xk+1 , Xk+2 , . . . , X2k ]T (7.4)

Then equation (7.1) can be written in terms of the state variables given by:

Ẋ = f (X1 , X2 , . . . , Xk , Xk+1 , Xk+2 , . . . , X2k , t)


(7.5)
= [f1 , f2 , . . . , f2k ]T

Equation (7.5) is called state equation. For equilibrium, Ẋ = 0 and (7.5)


becomes  
0 = f X̄1 , X̄2 , . . . , X̄k , X̄k+1 , X̄k+2 , . . . , X̄2k , t (7.6)
where X̄1 , X̄2 , . . . , X̄2k are the set of equilibrium state determined from (7.6).
To analyze the stability, the system is initially perturbed from the equi-
librium or the steady state with a small perturbation ∆X(0) in every coor-
dinates. Thus after the initial time, the state vector X(t) is then

X(t) = X̄ + ∆X(t) (7.7)

where

∆X(t) = [∆X1 (t), ∆X2 (t), . . . , ∆Xk (t), ∆Xk+1 (t), ∆Xk+2 (t), . . . , ∆X2k (t)]T
(7.8)
In (7.8), ∆X1 , ∆X2 , . . . , ∆X2k are small perturbation. Substitution of (7.7)
into the equations of motion (7.5) yields
˙ + ∆Ẋ(t) = f X̄ + ∆X , X̄ + ∆X , . . . , X̄ + ∆X , t
X̄ (7.9)
1 1 2 2 2k 2k

Equation (7.9) can be expanded using the Taylor’s series as


 
˙ + ∆Ẋ(t) ≈ f X̄ , X̄ , . . . , X̄ , t +

∂f
∆X(t) (7.10)
1 2 2k
∂X X=X̄

˙ = 0 and with the equilibrium condition (7.6), Equation (7.10)


Since X̄
becomes  
∂f
∆Ẋ(t) ≈ ∆X(t) = A∆X(t) (7.11)
∂X X=X̄
where  
A11 A12 ... A1,2k
   
∂f  A21 A22 ... A2,2k 
A≡ =
 .. .. .. .. 
 (7.12)
∂X X=X̄  . . . . 
A2k,1 A2k,2 . . . A2k,2k
7.2. STABILITY OF EQUILIBRIUM OR STEADY STATE 101

The component Aij in (7.12) is given by


 
∂fi
Aij = ; i = 1, 2, . . . , 2k; j = 1, 2, . . . , 2k (7.13)
∂Xj X=X̄

For a short notation, replace ∆X(t) in (7.11) with Y(t). Hence the pertur-
bation equation (7.11) becomes

Ẏ(t) = AY(t) (7.14)

To analyze the stability, we then solve for Y(t). First let the solution be

Y(t) = Ceλt (7.15)

Substituting (7.15) into (7.14) yields

λCeλt = ACeλt (7.16)

Or
(A − λI) Ceλt = 0 (7.17)
Thus the nontrivial solution of (7.17) is

|A − λI| = 0 (7.18)

From (7.18), there are 2k values of λ. The equilibrium or the steady state
will be unstable if either one of the following conditions is satisfied.

1. There exist real roots of λ and at least one of them is positive.

2. There exist complex roots of λ and at least one of them has a positive
real part.

Example 7.1 :
The bead is constrained to move along the circular ring as shown in Fig. 7.1.
If the ring is rotated about the vertical axis with a constant speed ω. De-
termine the steady states and analyze if each state is stable or unstable.
Solution
Kinematics:
The system has only one degree of freedom. From Fig. 7.1, let α be the
generalized coordinate. Hence the absolute velocity of the bead is

v = r α̇eθ + (rsinα) ωez (7.19)


102 CHAPTER 7. STABILITY ANALYSIS

ω constant

smooth
g

r eθ

er

Figure 7.1: Stability analysis of a bead steady motion

Kinetic energy T and potential energy V can be formulated as


1 1  
T = mv · v = m r 2 α̇2 + ω 2 r 2 sin2 α (7.20)
2 2
V = −mgrcosθ (7.21)
Lagrange equation is then
 
d ∂T ∂T ∂V
− + =0 (7.22)
dt ∂ α̇ ∂α ∂α
Formulate each term in (7.22) and substitute into the equation to get equa-
tion of motion:
mr 2 α̈ − mω 2 r 2 sinαcosα + mgrsinα = 0 (7.23)
Determine steady states:
Let’s define the state variables: X1 = α and X2 = α̇.
Hence the state equations are
Ẋ1 = X2
g (7.24)
Ẋ2 = ω 2 sinX1 cosX1 − sinX1
r
(7.24) can be put in the matrix form as
   
Ẋ1 X2
Ẋ = = g = f (X1 , X2 ) (7.25)
Ẋ2 ω sinX1 cosX1 − sinX1
2
r
7.2. STABILITY OF EQUILIBRIUM OR STEADY STATE 103

For the steady state, Ẋ = 0. Therefore, (7.25) becomes


   
0 X̄2
= g (7.26)
0 ω sinX̄1 cosX̄1 − sinX̄1
2
r

From (7.26), X̄2 = 0 and


 
g
sinX̄1 ω cosX̄1 −
2
=0 (7.27)
r
 
−1 g
There are two possible solutions for (7.27): X̄1 = 0 or X̄1 = cos .
ω2 r
Therefore the system has two steady states which are
     
(1) 0 (2) cos−1 g
ω2 r
X̄ = , X̄ = (7.28)
0 0

g
Note that the second steady state X̄(2) exists if and only if ≤ ω 2 .
r
Stability analysis
To analyze the stability of each steady state, the system is initially perturbed
with ∆X(0) from the steady state. After t > 0, the motion is described by
 
∆X1 (t)
X(t) = X̄ + (7.29)
∆X2 (t)

Substitute (7.29) into (7.25), and then linearize the equation. Let Y(t) =
∆X(t) for a short notation, the perturbation equation is therefore

Ẏ(t) = AY(t) (7.30)

where    
∂f1 ∂X2
A11 = = =0
∂X1 X=X̄ ∂X1 X=X̄
   
∂f1 ∂X2
A12 = = =1
∂X2 X=X̄ ∂X2 X=X̄
 
∂f2 g
A21 = = ω 2 cos2X̄1 − cosX̄1
∂X1 X=X̄ r
 
∂f2
A22 = =0
∂X2 X=X̄
104 CHAPTER 7. STABILITY ANALYSIS

Or the matrix A is
 
0 1
A= (7.31)
ω 2 cos2X̄1 − gr cosX̄1 0

Let the solution of (7.30) be Y(t) = Ceλt , λ can be obtained from the
characteristic equation: |A − λI| = 0, or
% %
% −λ %
% 1 %
% 2 % (7.32)
% ω cos2X̄1 − gr cosX̄1 −λ %

Equation (7.32) yields


 1
g 2
λ1,2 = ± ω cos2X̄1 − cosX̄1
2
(7.33)
r

For the 1st steady state, substitute X̄1 = 0 and X̄2 = 0 into (7.33) to get
 1
g 2
λ1,2 =± ω − 2
(7.34)
r
g
In (7.34), if ω 2 > then one of λ will be a positive real resulting in unstable
r
g
steady state. On the other hand if ω 2 ≤ then both λ will be conjugate
r
imaginary resulting in stable steady state.  
nd −1 g
For the 2 steady state, X̄1 = cos and X̄2 = 0. Equation
ω2r
(7.33) then becomes
! "1
λ1,2 = ± ω 2 cos2X̄1 − gr cosX̄1 2
   1
= ± ω 2 cos2 X̄1 − sin2 X̄1 − gr cosX̄1 2 (7.35)
 2 4 2 1
g −ω r
= ± ω2 r2
2
, g
r ≤ ω2

g
Since the 2nd steady state exists if and only if ≤ ω 2 , both λ in (7.35) are
r
conjugate imaginary. Thus this steady state is always stable.
Bibliography

[1] J. H. Ginsberg, Advanced Engineering Dynamics, Cambridge, 1998.

[2] D. T. Greenwood, Priciples of Dynamics, Prentice Hall, 1988.

[3] F. C. Moon, Applied Dynamics: With Applications to Multibody and


Mechatronic Systems, John Wiley & Sons, 1998.

105

You might also like