Anda di halaman 1dari 105

2103-617 Advanced Dynamics

Thitima Jintanawan

2
c
Copyright
2005
Thitima Jintanawan

Contents
1 Kinematics
1.1 Evolution of Kinematics . . . . . . . . . . . . . . . . . . . . .
1.2 Position Vector, Velocity, and Acceleration . . . . . . . . . . .
1.3 Angular Velocity . . . . . . . . . . . . . . . . . . . . . . . . .
1.4 Rate of Change of a Constant-Length Vector . . . . . . . . .
1.5 Moving Coordinate Systems . . . . . . . . . . . . . . . . . . .
1.6 Coordinate Transformation . . . . . . . . . . . . . . . . . . .
1.6.1 First set of Euler anglesprecession-nutation-spin ()
1.6.2 Second set of Euler anglesyaw-pitch-row () . . . .
1.7 Angular velocity related to Euler angles . . . . . . . . . . . .
1.8 A Finite Motion . . . . . . . . . . . . . . . . . . . . . . . . .
1.8.1 Transformation matrices for a nite rotation . . . . .
1.8.2 Transformation matrices for a general motion . . . . .

5
5
5
7
8
10
17
20
20
23
25
26
30

2 Linear and Angular Momentums


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

35
35
35
35
37
37
38
38
40
42

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

3 Dynamics of a Rigid Body


45
3.1 Newton-Euler Equations of a rigid body . . . . . . . . . . . . 45
3.2 Modied Eulers equations . . . . . . . . . . . . . . . . . . . . 49
3.3 Introduction to stability of a spin body . . . . . . . . . . . . 53
3

CONTENTS

4 Multi-Body Mechanical System


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

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

57
57
57
57
60
61
61
62

5 Principle of Virtual work


5.1 Virtual Displacement and Virtual Work .
5.2 Holonomic and Nonholonomic Constraints
5.3 Generalized Coordinates and Jacobian . .
5.4 Principle of Virtual Work . . . . . . . . .
5.5 DAlembert principle . . . . . . . . . . . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

69
69
69
71
75
76

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

6 Lagrange Mechanics
6.1 Kinetic Energy . . . . . . . . . . . . . . . . . . . . . . .
6.2 Potential Energy . . . . . . . . . . . . . . . . . . . . . .
6.3 Remarks on Properties of Generalized Coordinates for
System with Holonomic constraints . . . . . . . . . . . .
6.4 Derivation of Lagranges equations . . . . . . . . . . . .
6.5 Examples . . . . . . . . . . . . . . . . . . . . . . . . . .
6.6 Lagrange Multiplier . . . . . . . . . . . . . . . . . . . .

. . .
. . .
the
. . .
. . .
. . .
. . .

81
81
82
84
85
87
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 cylindrical coordinates, are introduced. Position vector, velocity and acceleration
of particles and rigid bodies are formulated using dierent reference coordinates. Each coordinate system is related to the other through the coordinate
transformation. For the 3-D transformation, two dierent sets of Euler angles: precession-nutation-spin and yaw-pitch-roll, are conventionally used.
Finally the transformation matrix used to describe a nite 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 approach
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. Lets


introduce a cartesian or rectangular coordinate system XY Z as shown in
5

CHAPTER 1. KINEMATICS

Z
rz

particle
moving path

r
k
j

ry

i
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
xed 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 cylindrical coordinates z, with their unit vectors e e ez . In Fig. 1.2, the position
vector r expressed in terms of e e ez is
r = e + zez

(1.3)

The absolute velocity v is dened as a time derivative of the position vector


r given by
dr
dt
= rx i + ry j + rz k
+ ze
z
= e
+ e

v =

(1.4)

1
An inertial or xed reference frame is the coordinate system whose origin O is xed
in space

1.3. ANGULAR VELOCITY

Z
ez

r
z
Y

Figure 1.2: A cylindrical coordinate system


The absolute acceleration a is dened as a time derivative of the velocity v
given by
dv
dt
= rx i + ry j + rz k




= 2 e + + 2 e + zez

a =

1.3

(1.5)

Angular Velocity

Figure 1.3 shows a rigid cylinder having a rotation about n axis. The absolute angular velocity of the rigid body is dened as
d
n
dt
= 1 e1 + 2 e2 + 3 e3

(1.6)

where 1 , 2 , and 3 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
3 2

= 3
0
1

2 1
0

(1.7)

CHAPTER 1. KINEMATICS

n
e3
e2

v
r

e1

A
(t)

Figure 1.3: Angular velocity


The velocity at point A in Fig. 1.3 is then
v = r

(1.8)

r
Equation (1.9) indicates that the cross product can be represented by the
matrix multiplication or

v1
0
3 2
r1

0
1 r2
v = v2 = 3
v3
2 1
0
r3

(1.9)

and r
Note that for the matrix multiplication in (1.9), components of
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 xed length vector c is given by the cross product of its rotation rate
and the vector c itself.
dc
=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 and simultaneously possesses a yaw maneuver (turn to left) with
Determine a relative velocity of point C on the horizontal
a rate of .
stabilizer at coordinates (b, a, 0), observed from the C.G. of the plane.

1.4. RATE OF CHANGE OF A CONSTANT-LENGTH VECTOR

ez

X
ex

G
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
xed length vector given in terms of the body coordinate system as
= bex + aey = [ b a 0 ]T
r(rel)
c

(1.11)

Hence the relative velocity of C is


(rel)
r
vc(rel) = r(rel)
c
c

(1.12)

z is the rotation rate or the angular velocity of the


y + e
where = e
reference coordinates moving with the body. can be written in a matrix
form as

0 0
0 0
Therefore

vc(rel)

0
b

T
a
=
0 0 = [ a b b]

0
0 0

As another example, the unit vectors ijk for any rotating system of
coordinates xyz is also the xed length vector. Hence the rate of change of
these ijk vectors can be determined from the same theorem as i = i,
j = j, and k = k, where is the angular velocity of such rotating
coordinate system xyz.

10

CHAPTER 1. KINEMATICS

z
Z

y
path of origin o
Y

Figure 1.5: Translating coordinate systems


z

y
o

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
y
o
path of origin o
x

Figure 1.7: Translating and rotating coordinate systems


body. Specically, such reference coordinate system is arranged such that
its origin o is xed to and translate with the bodys 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 and the position vector r
is
r =R+
(1.13)
Then the velocity v of the particle is given by
dr
dt
dR d
+
=
dt
dt
+ vr +
= R

v =

(1.14)

is the velocity of the origin o of the reference frame


xyz
 
d
,
and 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 velocity observed by the observer moving (both translating and rotating) with
xyz, whereas ddt is the relative velocity observed by the observer who only
translates but not rotates with xyz. in (1.14) is hence the dierence
between these two relative velocities. If is

=
In (1.14), R

dR
dt

= 1 e1 + 2 e2 + 3 e3

(1.15)

12

CHAPTER 1. KINEMATICS

z
e3
Z

e2

O'
x

path of moving
reference frame

e1

path of particle
r

k
i

Figure 1.8: Moving coordinate systems


Then

vr

d
dt

= 1 e1 + 2 e2 + 3 e3

(1.16)

rel

The absolute acceleration a of the particle is then


dv
dt
+ ar + vr + d + d
= R
dt
dt

a =

(1.17)

From (1.14)
d
= vr +
dt

(1.18)

+ ar + + + 2 vr
a=R

(1.19)

Plug (1.18) into (1.17) yields

where ar = 1 e1 + 2 e2 + 3 e3 .
We can describe the physical meaning of each term in (1.19) as follows.
is the acceleration of the origin o of the moving reference frame xyz.
R
ar is the relative acceleration of the particle as observed in the moving
reference frame xyz.

1.5. MOVING COORDINATE SYSTEMS

13

is a centripetal acceleration, or the correction term for the


local position vector considering that the observer rotates with the
moving reference frame.
is another correction term for the angular acceleration vector
of the moving reference frame.
2 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 about the body xed axis
and angular
e3 The solar panel arm rotates about the e2 -axis with a rate ,

acceleration = 0. The panel arm also moving along the radial direction er
with a steady rate s = . 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
e1 e2 e3 = e3 . Choose [er e 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 , = r er + e + 2 e2 = (s(t) + c)er
2
= e3 + e
Note that R and are the xed-length vectors with constant magnitudes.
From (1.19), the absolute acceleration of point P is
+ ar + + + 2 vr
ap = R

(1.20)

where
= e3 R = be2 , R
= e3 R
= b2 e1
R = be1 , R
vr = r er + e + 2 e2 = s(t)e

r = er
ar = r er + e + 2 e2 = s(t)er = 0
1
= e3 = e
Components in (1.20) are now expressed in terms of two dierent coordinate
systems [e1 e2 e3 ] and [er e 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

s(t)

+c

e2
e1

e2
b

Figure 1.9: A satellite

e3

er

ur
u
e1

u1

u3
e2

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

1.5. MOVING COORDINATE SYSTEMS

15

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


vector u as


u=

ur
u

cos sin
sin cos



u3
u1

=T

u3
u1

The reader can prove that T is an orthogonal matrix or T1 = TT . As


a result


u3
u1

cos sin
sin cos

1 

ur
u

=T

ur
u

=T

ur
u

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


terms of [e1 e2 e3 ] components as
vr = er = (cose3 + sine1 ) = [ sin, 0, cos ]T
= [ 0 ]T

=
0 0

0 0
= [ 0 0 ]T

0
0
0

= 0

0

0 0
= (s(t)+c)er = (s(t)+c) (cose3 + sine1 ) = (s(t)+c)[ sin, 0, cos ]T
= [ b2 0 0 ]T
R
Plug these terms into (1.20), we obtain ap in terms of the rotating system
of coordinates [e1 e2 e3 ] as
ap = [ b2 0 0 ]T + 0

0
0

0 0
0
+(s(t) + c)

0 0 0

0
0
0
sin

+(s(t) + c) 0
0
0
cos
0 0

0
sin

+2
0 0 0
cos
0 0

sin

0 0
cos
0
(1.21)

16

CHAPTER 1. KINEMATICS

z1

z2

y2

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 e e2 ].
Example 1.3 :
Figure 1.11 shows a ventilator mounted on a rotating base. The base has
an oscillatory motion with = 0 sint. The rotor spins with a constant
angular velocity in the direction shown. Its center of gravity G is oset
by from the axis of rotation. Determine:
1. components of the absolute angular velocity of the rotor along the
system of coordinates xed 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
z2
x
z

cos sin
sin cos

cost sint
sint cost





y1
z1
x2
z2

1.6. COORDINATE TRANSFORMATION


Z, z1

17
y1

z2

z
t

x1
z1

z2

y2, y
t

x2
y2

x1, x2

y1

Figure 1.12: Coordinate Systems


Absolute angular velocity of the rotor is then
=
=
=
=
=

e
z1 + ey2
(sine

y2 + cosez2 ) + ey2
(sin

+ )ey2 + cose

z2
(sin

+ )ey + cos(sinte

x + costez )

cossinte

x + ( + sin)e
y + coscoste
z

The position vector of G is


rG = Ley2 + ez
= Ley + ez
= [ 0 L ]T
Absolute velocity of G is
rG
0

L
=

r G =

1.6

Coordinate Transformation

In Section 1.5, we simply transform the coordinates in two dimensional


(2-D) space. Now lets consider a general 3-D coordinate transformation.
Specically, 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
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
rZ
rz

(1.25)

1.6. COORDINATE TRANSFORMATION

19

i
Ji

iJ
iY

Figure 1.14: components of the unit vectors


where C is a matrix of directional cosines so called a coordinate transformation matrix given by

cos iI cos jI cos kI

C = cos iJ cos jJ cos kJ


cos iK cos jK cos kK

(1.26)

Note that C is the orthogonal matrix where C1 = CT . This yields

rx
rX

T
ry = C rY
rz
rZ

(1.27)

Q: Are all nine components of C independent?


To answer this question, lets consider Fig. 1.14, showing the following
relations.
|iY |
= |iY |
(1.28)
cos iJ =
|i|
With similar expressions for the other axes, we come up with the following
6 relationships
|iX |2 + |iY |2 + |iZ |2
|jX |2 + |jY |2 + |jZ |2
|kX |2 + |kY |2 + |kZ |2
|Ix |2 + |Iy |2 + |Iz |2
|Jx |2 + |Jy |2 + |Jz |2
|Kx |2 + |Ky |2 + |Kz |2

= cos2  iI + cos2 
= cos2  jI + cos2 
= cos2  kI + cos2 
=
cos2  Ii + cos2 
= cos2  Ji + cos2 
= cos2  Ki + cos2 

iJ + cos2 
jJ + cos2 
kJ + cos2 
Ij + cos2 
Jj + cos2 
Kj + cos2 

iK
jK
kK
Ik
Jk
Kk

=
=
=
=
=
=

1
1
1
1
1
1

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


independent components of C. Specically, 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 anglesprecession-nutation-spin ()

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 (t) to get x y  z  or x y  Z
Nutation: rotation about x axis by (t) to get x y  z  or x y  z 
Spin: rotation about z  axis by (t) to get xyz or xyz 
The coordinate transformations are then

rX
cos sin 0
rx
rx

rY = sin cos 0 ry = C1 ry


rZ
rz 
rz 
0
0
1

rx
1
0
0
rx
rx



r
=
r
=
C
0
cos
sin
y
y
2 ry 
rz 
rz 
rz 
0 sin cos

(1.29)

(1.30)

rx
cos sin 0
rx
rx

ry = sin cos 0 ry = C3 ry


rz 
rz
rz
0
0
1

(1.31)

Combine (1.29)-(1.31), therefore

rX
rx

rY = C1 C2 C3 ry
rZ
rz

1.6.2

(1.32)

Second set of Euler anglesyaw-pitch-row ()

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 (t) to get x y  z  or x y  Z

1.6. COORDINATE TRANSFORMATION

21

X
x
z

x x
z

y
y

x
x

Figure 1.15: First set of Eulers angles and sequence of rotation

22

CHAPTER 1. KINEMATICS

Z z

Y
x
z

y y

X
z

x x

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 (t) to get x y  z  or x y  z 


Roll: rotation about x axis by (t) to get xyz or x yz
The coordinate transformations are then

rX
cos sin 0
rx
rx

rY = sin cos 0 ry = [R ] ry


rZ
rz 
rz 
0
0
1

rx
cos 0 sin
rx
rx

0
1
0 ry = [R ] ry
ry =
rz 
rz 
rz 
sin 0 cos

(1.34)

rx
1
0
0
rx
rx

ry = 0 cos sin ry = [R ] ry


rz 
rz
rz
0 sin cos

(1.35)

Therefore

rX
rx

rY = [R ] [R ] [R ] ry
rZ
rz

1.7

(1.33)

(1.36)

Angular velocity related to Euler angles

For the rst set of Euler angles, the absolute angular velocity of xyz
coordinates is given by
z
+ e
x + e
= k
x ex + y ey + z ez

(1.37)

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

x
sinsin cos 0



y = sincos sin 0
z
cos
0
1

(1.38)

Similarly, for the second set of Euler angles, the absolute angular velocity
of xyz coordinates is given by
x
+ e
y + e
= k
= x ex + y ey + z ez

(1.39)

24

CHAPTER 1. KINEMATICS

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

x
1
0
sin

=
0
cos
cossin
y

z
0 sin coscos

(1.40)

(1.38) and (1.40) relate the Euler angles, the rotation that measured in real
applications, with the components of the angular velocity, x , y and z , in
the reference coordinate system.
Example 1.4 :
A submarine shown in Fig. 1.17 undergoes a yaw rate = Acost and a
pitch rate = Bsint. 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 bodyxed rotating system of coordinates. The velocity of the bow observed from
the submarine C.G. is then
v =

(1.41)

= Lex

(1.42)

where

and is the angular velocity of the body or the angular velocity of the xyz
coordinates given by
+ e
y
(1.43)
= k

1.8. A FINITE MOTION

25

A'
u

Figure 1.18: A nite motion of a rigid body


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

0
x
1
0
sin



= y = 0 cos cossin
z
0 sin coscos

(1.44)

Note that, in this case, the submarine performs only pitch and yaw rotations
but no row. Neglecting the higher order terms, is therefore

= sine
x + ey + cosez

(1.45)

Substitution of (1.42) and (1.45) into (1.41) yields

v = Lcose
y Lez

(1.46)

From given = Acost and = Bsint, and if (0) = (0) = 0, then


B
(t) = A
sint and (t) = cost. Substitution of these conditions into
(1.46) yields


B
cost ey LBsintez
v = ALcostcos

(1.47)

B
For the small value of B
, cos cost 1. In addition if A = B, the
velocity vector v = AL(cos tey + sin tez ) 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 nite rotation about this point

26

CHAPTER 1. KINEMATICS

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


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

1.8.1

Transformation matrices for a nite rotation

Dene 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
(1.48)
rp = Trp
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 = T1 r

(1.50)

3. The length of r is unchanged, hence


 T 
r

r r = rT r = r r = r

(1.51)

Or
rT r =

  T 
r r

(1.52)

= (Tr)T Tr
= rT TT Tr
Hence
TT T = I

(1.53)

(1.53) indicates that T1 = TT or T is the orthogonal matrix.


The transformation T for a rotation about Z-, Y -, and X-axes can be determined 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
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
cos sin 0
x
x


y = sin cos 0 y = T1 y
z
0
0
1
z
z
The transformation matrix T1 in this case is

(1.54)

cos sin 0

T1 = sin cos 0
0
0
1

(1.55)

2. Rotation about Y -axis with


From Figure 1.20 we obtain the transformation matrix T2 for a rotation about Y -axis as

cos 0 sin

0
1
0
T2 =
sin 0 cos

(1.56)

3. Rotation about X-axis with


From Figure 1.21, the transformation matrix T3 for a rotation about
X-axis is

1
0
0

(1.57)
T3 = 0 cos sin
0 sin cos

28

CHAPTER 1. KINEMATICS

Figure 1.20: Finite rotation about Y -axis

Figure 1.21: Finite rotation about X-axis

1.8. A FINITE MOTION

29

Figure 1.22: A falling box


In general the transformation matrices do not commute; i.e., T1 T2 = T2 T1 .
However, for the innitesimal angular displacement, cos 1, sin and
so on. Also z t, y t, and x t. In this case T1 , T2 , and
T3 do commute. If the rigid body has an innitesimal 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
z t y t
x(t)

1
x t y(t)
r(t + t) = z t
1
y t x t
z(t)

(1.59)

Therefore the velocity v is given by

0
z y
x(t)
r(t + t) r(t)

(1.60)
= z
0
x y(t) r
v = lim
t0
t
y x
0
z(t)
can be represented
From (1.60), it is proved that the angular velocity
in a matrix form as previously introduced in (1.7).
The transformation matrix for a nite rotation is useful for a computer graphic programming simulating the dynamics of rigid-body motion
as shown in the following example.
Example 1.5 :

30

CHAPTER 1. KINEMATICS

O
R

b
mg

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-le 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


mgLsin( + 0 ) C = Io

Mo = H o ];

(1.61)

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


the hinge and Io is the mass moment of inertia about o. To solve (1.61),
lets dene state variables as x1 = and x2 = Then (1.61) can be written
in state form as


x 1
x 2

x2
esin(x1 + 0 ) cx2

(1.62)

C
where e = mgL
Io and c = Io . (1.62) together with the transformation matrix
for the nite 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 nite rotation . The position vector r

1.8. A FINITE MOTION


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

31

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

0.05

0.1

0.15

0.2

0.25

0.3

Figure 1.25: Simulation of the falling box

ez
A

r
ex
k

A
j
i

Figure 1.26: Finite motion

1.8. A FINITE MOTION

33

describing the nite rotation of the rigid body is then


r = u + 

(1.63)

= u + A
where A is the transformation matrix of the rotation relating and  . In
addition r = [ x y z ]T and u = [ ux uy uz ]T expressed in the inertial
reference coordinates xyz, and = [ x y z ]T expressed in the local
coordinate system ex ey ez . Substituting the component vectors into (1.64),
the nite motion in Fig. 1.26 is governed by

x
ux
cos sin 0
x

y = uy + sin cos 0 y
uz
z
z
0
0
1

(1.64)

If r and are expanded as r = [ x y z 1 ]T and = [ x y z 1 ]T ,


(1.64) can be rewritten as
r41 = T44 41

(1.65)

where T is the transformation matrix for a general nite motion given by

T=

A33

|
ux
|
uy
|
uz
|
0
|
1

For a nite translation, the transformation matrix is simply

T1 =

1
0
0
|
ux
0
1
0
|
uy
0
0
1
|
uz
|
0
0
0
|
1

For a nite rotation, the transformation matrix is simply

A33

T2 =

0
Note that T = T1 T2 .

|
0
|
0
|
0
|
0
|
1

34

CHAPTER 1. KINEMATICS

Chapter 2

Linear and Angular


Momentums
This chapter is organized into three parts: 1) dynamics of a system of particles; 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

where mi is a mass of the i-th particle and


1, 2, . . . , n.

2.1.2

(2.1)
is the sum over i, for i =

First moment of mass

The rst moment of the total mass about its C.G. is the sum of the rst
moment of each mass given by:
rc M =

35

ri mi

(2.2)

36

CHAPTER 2. LINEAR AND ANGULAR MOMENTUMS

Fi

mi
vi

mN
FN

Z
ri
F1

rc

m1

c
Y

F3

m3
F2
m2

Figure 2.1: A system of particle

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


rc =

1 
ri mi
M

(2.3)

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

(2.4)

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


i mi =
[r rc ] mi
 i

=
ri mi rc mi
= rc M rc M
= 0

(2.5)

Equation (2.5) indicates that sum of the rst moment of each mass about
the systems C.G. is zero.

2.1. DYNAMICS OF A SYSTEM OF PARTICLES: A REVIEW

2.1.3

37

Linear momentum

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




mi vi
dri
mi
dt

d 
mi ri
dt
d
(M rc )
dt
M vc


=
=
=
=

2.1.4

(2.6)

Angular momentum

An angular momentum is dened as the rst moment of the linear momentum. The angular momentum of the system of particles about the origin o
is then given by

ri mi vi
(2.7)
Ho
The angular momentum of the system of particles about the systems C.G.
is dened as

i mi i
(2.8)
Hc
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




[(i + rc ) mi vi ]
Ho =


i mi vi + rc mi vi
=

(2.10)

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


rc

mi vi = rc M vc

(2.11)

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




i mi vi =
mi (rc + i )
 i

mi r c + i mi i
=
 i
=
(mi i ) r c + Hc
= 0 + Hc

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

(2.12)

38

CHAPTER 2. LINEAR AND ANGULAR MOMENTUMS

2.1.5

Moment of force

The moment due to all applied forces about o is


Mo =

2.1.6

ri Fi

(2.13)

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 Newtons 2nd law to each i-th particle, we obtain
mi v i = Fi +

i = j

fij ;

(2.14)

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 v i =

Fi +



Since fij = fji , therefore


i

i = j

fij ;

(2.15)

fij = 0. Hence (2.15) becomes


dP
= M v c =
Fi
dt
i

(2.16)

(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
Lets take the rst moment of (2.14) about o and sum over all particles:

i

(ri mi v i ) =


i

(ri Fi ) +


i

ri


j

fij ;

i = j

(2.17)

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


(ri mi v i ) =

ri mi

dvi
dt

d 
ri mi vi
=
dt i
o
= H

The rst term on the right of (2.17) is

(2.18)

(ri Fi ) Mo , and the second

term on the right is zero as shown in the following proof.


Lets 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



i

ri

fij = 0

(2.20)

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


dHo
o = Mo
=H
dt

(2.21)

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 systems C.G.. First, dierentiate (2.9) with time:
d
d
d
Ho = Hc + (rc M vc )
dt
dt
dt

(2.22)

From (2.22), the rst term in (2.22) is rewritten as


d
Ho = Mo
dt

=
r Fi
 i
=
(r + i ) Fi
 c

=
rc Fi + i Fi

(2.23)

40

CHAPTER 2. LINEAR AND ANGULAR MOMENTUMS

e3

e2

dm

A
e1
rA

o
X

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
= vc M vc + rc Fi

= 0 + rc Fi

(2.24)

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




Or
c=
H

i Fi =


d
Hc
dt

i Fi = Mc

(2.25)

(2.26)

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


we can dene 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 . 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

r vdm

Ho =

(2.27)

where r = rA + is a position vector from the xed origin o to the dierential


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. , then
(2.28)
v = r = vA +
Note that the velocity of the dierential mass dm relative to e1 e2 e3 is zero
because: 1) the rigid body has no deformation and 2) e1 e2 e3 rotates synchronously with the body. Substitution of (2.28) into (2.27) yields


[(rA + ) (vA + )] dm

Ho =

= rA vA



dm +

+rA

dm vA

dm +

( ) dm

= rA mvA + mc vA + rA [ mc ] +

( ) dm

(2.29)
of
where m = dm is the mass of the rigid body and c is the position
1 
dm.
the bodys C.G. measured with respect to A and given by c = m
Furthermore, the rotation of a rigid body can be considered as two dierent
cases: pure rotation and general motion (combined rotation and translation).


1. Pure rotation about xed point o


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


Ho =

( ) dm

(2.30)

2. General motion
In this case if point A in Figure 2.2 is xed at the bodys C.G, i.e.
point c. Hence rA = rc and c = 0. (2.29) then becomes
Ho = rc mvc +

( ) dm

(2.31)

42

CHAPTER 2. LINEAR AND ANGULAR MOMENTUMS


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

( ) dm

(2.32)

Hence
Ho = rc mvc + Hc

2.3

(2.33)

Mass Moment of Inertia

Due to a constant geometric property of the rigid body, the angular momentum can be more simplied as follows. First it is noted that the angular
momentum about o (2.30), in case of pure rotation, and the angular momentum 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 generally rewrite both equations as


H=

( ) 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=

( ) ( )dm

(2.35)

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


coordinate system as follows


x y z

T

x y z

T

Next we will rewrite (2.35) in terms of its components. Lets consider (2.35)
term by term. The rst term is


( )dm =




( ) [] dm

2 0 0

=
0 2 0 dm
0 0 2


2 0 0

= 0 2 0 dm
2
0 0

(2.36)

2.3. MASS MOMENT OF INERTIA

43

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


(2.35) is

( )dm =

x
x

x y z y y dm
z z
x

(xx + yy + zz ) y dm
z

2
x x + xyy + xzz

xyx + y 2 y + yzz dm
xyx + yzy+z 2 z

x2 xy xz
x

2
yz y dm
xy y
xz yz z 2 z

x2 xy xz

xy y 2 yz dm()
xz yz z 2


(2.37)

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

H =



y2 + z2

xy

I11

= I21
I31

xy
xz

2
2
yz dm
x +z
yz
x2 + y 2

xz
I12 I13

I22 I23 = I
I32 I33

(2.38)

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


along diagonal are called moments of inertia given by
 

I11 =
I22 =
I33 =

 
 

y 2 + z 2 dm


x2 + z 2 dm


x2 + y 2 dm

(2.39)

44

CHAPTER 2. LINEAR AND ANGULAR MOMENTUMS

and the o-diagonal components so called cross product of inertia are given
by

I12 = I21 = xydm
I13 = I31 =
I23 = I32 =




(2.40)

xzdm
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 simplied as
H = I1 1 e1 + I2 2 e2 + I3 3 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 o-diagonal terms
or the cross-product of inertia are zero.
4. The parallel axes theorem states that
(c)

Ikk = Ikk + m2k ,

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 dierent 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 ,
the governing equation is

o
Mo = H
(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
xed at o, that rotates with the body with the same angular velocity .
Hence, the angular momentum about o can be simplied as
Ho = I o

(3.2)

where Io is the constant matrix of moments of inertia about o whose components are along {e1 e2 e3 } axes.
For a general motion of a rigid body, the equations governing both translation and rotation are

F = mv c
(3.3)
and

c
Mc = H

where point c is the C.G. of the rigid body.


45

(3.4)

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 Newtons 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 xed at the C.G. of the body and its coordinates rotate
with the body. If the rigid body has the angular velocity , then
Hc = I c

(3.5)

where Ic is the constant matrix of moments of inertia about c whose components 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 = I1o 1 e1 + I2o 2 e2 + I3o 3 e3

(3.6)

Hc = Ic = I1c 1 e1 + I2c 2 e2 + I3c 3 e3

(3.7)

and
Hence (3.1) and (3.4) can be more simplied as


M1o
I1o 1
0
3 2
I1o 1

Mo = M2o = I2o 2 + 3
0
1 I2o 2
M3o
I3o 3
2 1
0
I3o 3
(3.8)

and
M1c
I1c 1
0
3 2
I1c 1


Mc = M2c = I2c 2 + 3
0
1 I2c 2
M3c
I3c 3
2 1
0
I3c 3
(3.9)
Or they can be written in a scalar form as

and

M1o = I1o 1 + (I3o I2o ) 2 3


M2o = I2o 2 + (I1o I3o ) 1 3
M3o = I3o 3 + (I2o I1o ) 1 2

(3.10)

M1c = I1c 1 + (I3c I2c ) 2 3


M2c = I2c 2 + (I1c I3c ) 1 3
M3c = I3c 3 + (I2c I1c ) 1 2

(3.11)

Equation sets (3.10) and (3.11) are called Eulers equations.

3.1. NEWTON-EULER EQUATIONS OF A RIGID BODY

47

u(t)
massless cart

ey
ex
g

m, L

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 oor 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 (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

(3.12)
vc = u(t)e
x + e
2
L
L
(t)ex + e 2 er
(3.13)
ac = v c = u
2
2
With the free body diagram (FBD) shown in Figure 3.2, we set NewtonEuler equations as

[ F = mv c ];


and [

L
L
(t)ex + e 2 er
Fr er + F e mgey = m u
2
2

(3.14)

Mc = Ic ];

L
= Ic
2

(3.15)

48

CHAPTER 3. DYNAMICS OF A RIGID BODY

Fr
ey

ex

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 , F , and , 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 F and
reduce the Newton-Euler equations to only one dierential equation.
Figure 3.2 shows the two coordinate systems with the coordinate transformation given by
ex = siner + cose
(3.16)
ey = coser + sine
To eliminate F , we substitute (3.15) into (3.14) and transform all coordinates to {er e } using (3.16). Then (3.14) can be expressed in scalar form
as:
r-component:
m 2

L
+ mgcos = Fr + m
usin
2

(3.17)

-component:


mL2
Ic +
4

mL
mgL
sin =
u
cos
+
2
2

(3.18)

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 integration using Matlab. Otherwise if only small oscillation is interested, we can

3.2. MODIFIED EULERS EQUATIONS

49

e2
dm

e3

C.G. e
1

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 F are obtained from (3.15) and (3.17) as
L
usin
Fr = m 2 + mgcos m
2

(3.19)

and
F =

3.2

2Ic

(3.20)

Modied Eulers equations

A set of modied Eulers equations is used in the case of the symmetricshape rigid body which spins about its symmetry axis with a constant speed,
as shown in Figure 3.3. To formulate the Modied Eulers equations, two
conditions are dened.
1. The rigid body spins about the symmetry axis with a constant speed
o .
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 xed at
the point of rotation for the case of pure rotation, and is xed at the

50

CHAPTER 3. DYNAMICS OF A RIGID BODY


bodys 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
= 1 e1 + 2 e2 + 3 e3
The angular velocity of the rigid body is then
b = + o e3

According to the conditions above, the modied Eulers equations become

and

1 + (I3o Io ) 2 3 + I3o o 2
M1o = Io
2 + (Io I3o ) 3 1 I3o o 1
M2o = Io
M3o = I3o 3

(3.21)

1 + (I3c Ic ) 2 3 + I3c o 2
M1c = Ic
2 + (Ic I3c ) 3 1 I3c o 1
M2c = Ic
3
M3c = I3c

(3.22)

Equation sets (3.21) and (3.22) are for the cases of pure rotation and general
motion, respectively. The derivation of the modied Eulers 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 =
=
=
=
=

dm
( )


d



dt
+ dm
rel

( [(o e3 ) + ]) dm

( (o e3 + ) ) dm

(
b ) dm

Ic 0 0
1

2
= I b = 0 Ic 0

3 + o
0 0 I3c

(3.23)

Then

c
H


= I 2 + Hc

1
Ic
0
3 2
Ic 1

2
= Ic
0
1
Ic 2
+ 3

0
I
(
+

)
I3c 3
2
1
3c
3
o

(3.24)

3.2. MODIFIED EULERS EQUATIONS

51

e3

L
k
e1

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
and the nutation angle is also
precession rate and constant spin rate ,
constant.
Method1: Direct approach In Figure 3.4, the angular velocity of the
reference coordinate system e1 e2 e3 is
+ e
2
e1 e2 e3 = k

(3.25)

Also the angular velocity of the body is


b
=
=
=

1 e1 + 2 e2 + 3 e3
+ e
2 + e
3
k
2 + e

(cose3 + sine1 ) + e
 3

sine
1 + e2 + + cos e3

(3.26)

and 3 = + cos.

2 = ,
Since e3 is the
Hence 1 = sin,
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 b . Or
Ho = I1 1 e1 + I1 2 e2 + I3 3 e3


= I1 sine
1 + I1 e2 + I3 + cos e3

(3.27)

52

CHAPTER 3. DYNAMICS OF A RIGID BODY


For a steady motion, is constant or = 0, and 1 = 2 = 3 = 0.
The angular momentum is then


Ho = I1 sine
1 + I3 + cos 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
Ho
o = k
H

(3.29)

Substituting (3.28) into (3.29) and performing a matrix operation yield




o = (I1 I3 ) 2 sincos I3 sin


e2
H


(3.30)

o , the moment sum


Mo = H
From the law of angular momentum

Mo about o, in this case, is due to only the gravitation force, given
by

Mo = Le3 mgk = mgLsine2
(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 . Substitution of (3.30)
and (3.31) into the law of angular momentum yields

mgLsin = (I1 I3 ) 2 sincos I3 sin

(3.32)

For sin = 0, we get the equation governing a steady precession of the


top as
(3.33)
mgL = I3 + (I3 I1 ) 2 cos
From (3.33), with a given we can determine the relation between the
For example, if = equation
precession rate and the spin rate .
2
(3.33) becomes
mgL
(3.34)
=
I3
Method 2: modied Eulers equations The modied Eulers equations
are presented again as follows
1 + (I3o Io ) 2 3 + I3o o 2
M1o = Io
2 + (Io I3o ) 3 1 I3o o 1
M2o = Io
3
M3o = I3o

(3.35)

3.3. INTRODUCTION TO STABILITY OF A SPIN BODY

53

Figure 3.5: Stability of a spin plate

1 = 2 =
2 = 0, and 3 = cos.
Also
In (3.35), 1 = sin,

0 because of a steady motion. Furthermore, the spin rate o = .


Substitution of these terms into (3.35) yields

M2o = mgLsin = (I1 I3 ) 2 sincos I3 sin

(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, lets rst formulate the Eulers equations for

54

CHAPTER 3. DYNAMICS OF A RIGID BODY

the spin plate as follows:


M1c = I1c 1 + (I3c I2c ) 2 3
M2c = I2c 2 + (I1c I3c ) 1 3
M3c = I3c 3 + (I2c I1c ) 1 2

(3.37)

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. Lets assume that the plate has a steady spin about the
axis 1 with a constant speed 0 . (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 1 (t), 2 (t), and 3 (t), respectively, about all
principal axes. Hence the angular velocities in each direction are
1 (t) = 0 + 1 (t)
2 (t) = 2 (t)
3 (t) = 3 (t)

(3.38)

Substitution (3.38) into (3.37) and neglecting the higher order terms, such
as 1 2 , 2 3 , etc., yield
= 0
I1c 1
I2c 2 + (I1c I3c ) 0 3 = 0
I3c 3 + (I2c I1c ) 0 2 = 0

(3.39)

The rst row of (3.39) implies that 1 (t) is constant. In addition, the last
two rows of (3.39) can be written in a matrix form as


2 (t)
3 (t)

0
(I2c I1c )0
I3c

(I1c I3c )0
I2c



2 (t)
3 (t)

0
0

(3.40)

or

(t)
+ K(t) = 0

(3.41)

To solve (3.40), assume the solution as the following form




(t) =

2 (t)
3 (t)

a
b

et

(3.42)

Substitution (3.42) into (3.41) yields




[I + K]

a
b


t

0
0

(3.43)

3.3. INTRODUCTION TO STABILITY OF A SPIN BODY

55

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


The characteristic roots can be solved as
2 =

(I1c I3c ) (I2c I1c ) 02


I2c I3c

(3.44)

There are two roots of which are




1,2

(I1c I3c ) (I2c I1c ) 02


=
I2c I3c

1
2

(3.45)

With two roots, the solution (3.42) is then




2
3

a1
b1


1 t

a2
b2

e2 t

(3.46)

To analyze the stability from the values of , we can divide 2 into two cases
as follows
Case I: (2 0) In this case, 1,2 are positive and negative imaginary parts
and the rotation are marginally stable. Specically, 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: (2 > 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 gure 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 (coordinates) 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 connecting 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

z
z

ry
x

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, x = 0 and z = 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 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 equations, 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 , , and are the DOFs
used to describe the rotation.
Example 3: rolling sphere Consider the spherical ball rolls without slipping as shown in Fig. 4.3. The rst 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 = rcx y a = 0

4.3. CONSTRAINT EQUATIONS

59

Rz

Ry

Rx

Figure 4.2: Ball and socket

z
a
y
x

rcz
A

rcy

Figure 4.3: A rolling sphere

rcx

60

CHAPTER 4. MULTI-BODY MECHANICAL SYSTEM


(vAy )rel = rcy + x a = 0
Or the velocities of the C.G. are then
vcx = rcx = y a
vcy = rcy = x 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 lets 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 , , 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 =6N

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



C is number of all constraint equations.

4.4

Classication 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 classied as holonomic constraints. In addition, the holonomic constraints 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 classied as nonholonomic constraint, e.g. example 3 in Section 4.3: rolling sphere.

4.5. NUMBER OF DOF VS. DRIVING FORCES

61

z
z

Z
x

Y
rY
X

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

C L
M =6N
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. Dene 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-Eulers 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-Eulers
equations
4. Eliminate all unknown forces to obtain equations of motion (number 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 Figure 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 (in rad/s) as Md = M0 M , 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: (0) = 0, (0)


= 0, (0) = 0, and (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 and 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,
1
1 = k

(4.1)

4.7. EXAMPLE PROBLEM: DYNAMICS OF TWO-LINK ARMS

63

z2
MAy
A

MAz

y1

RAy

x1

y2

RAx
RAz
c

Z, z1

FBD of link 2

z2

mg

Z, z1
RAz

A
y1
m, l

rG

y;
;
y

Link 1

Md

y2

MAy

RAx

RAy
MAz

Link 2

Md

Figure 4.5: Two-link arms

MCY

RCY
RCX
MCX

RCZ

FBD of link 1

64

CHAPTER 4. MULTI-BODY MECHANICAL SYSTEM


2
1 + i
2 = k
2 + sinj
= i

2 + cosk
2

(4.2)

l
l

= a sin
i2 + j
2
2
2

(4.3)

vG2

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


v G2 =

a
2l
sin 2l cos
i2


l
l 2
2
+ a cos sincos j2
2

l 2
2

(4.4)

+ a 2 sin + 2l 2 sin2 k2

Kinetic analysis
Figure 4.5 shows the free body diagram of both links. First lets consider
link 2. The Newtons equation governing the translation of link 2 is
mv G2 = Fx2 i2 + Fy2 j2 + Fz2 k2

(4.5)

where the resultant forces are determined from the free body diagram as
Fx2 = RAx ,

Fy2 = RAy mgsin,

Fz2 = RAz mgcos

(4.6)

Eulers equations governing the rotation of link 2 are


M1c = I1c 1 + (I3c I2c ) 2 3
M2c = I2c 2 + (I1c I3c ) 1 3
M3c = I3c 3 + (I2c I1c ) 1 2

x2 :
y2 :
z2 :
where

I1c = I2c =

1 = ,

ml2
,
12

2 = sin,

(4.7)

I3c = 0

(4.8)

3 = cos

(4.9)

l
l
M2c = RAx + MAy ,
M3c = MAz
(4.10)
M1c = RAy ,
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
sin cos

= RAx
m a
2
2


(4.11)

l
l
a 2 cos 2 sincos = RAy mgsin
2
2

(4.12)

4.7. EXAMPLE PROBLEM: DYNAMICS OF TWO-LINK ARMS




65

l
l 2
+ a 2 sin + 2 sin2 = RAz mgcos
2
2
RAy

ml2 ml2 2
l
=

sincos
2
12
12


l
ml2 

RAx + MAy =

sin + 2 cos
2
12
MAz = 0

(4.13)
(4.14)
(4.15)
(4.16)

Now lets 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 Eulers equation in z1 -direction which is
Md MAz cos MAy sin + RAx a = 0

z1 :
Or

MAy =

Md
RAx a
MAz cot +
sin
sin

(4.17)

(4.18)

Plug (4.18) and (4.16) into (4.15) to obtain



Md
RAx a
ml2 
l

+
=

sin + 2 cos
RAx +
2 sin
sin
12

(4.19)

Then plug (4.11) into (4.19) to eliminate RAx and rearrange the equation
as get
2

+ ml3
sin2 + mal
sin +
ma2
mal
=0
+ 2 cos (M0 M )

5
2 sincos

12 ml

(4.20)

To eliminate RAy , plug (4.14) into (4.12) and rearrange the equation as
2 2 2
l l sincos a 2 cos + gsin = 0
3
3

(4.21)

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)
x2 = , and
in state form. First, lets dene the state variables x1 = ,
By substituting the state variables into (4.20) and (4.21), the
x3 = .
equations of motion can be put into in the state form as follows.

x 1
f1

x = x 2 = f (x1 , x2 , x3 ) = f2
x 3
f3

(4.22)

66

CHAPTER 4. MULTI-BODY MECHANICAL SYSTEM

Time reponses with zero initial conditions


(solid) and (dash); rad/s

2.5
2
1.5

1
0.5
0

-0.5

5
time (s)

10

5
time (s)

10

(degree)

5
4
3
2
1
0

Figure 4.6: Time responses of the two link arms


where
f1 =

M0 M x1 (mal/2)x1 x3 cosx2 (5ml2 /12)x1 x3 sinx2 cosx2


ma2 + (ml2 /3)sin2 x2 + malsinx2
f2 = x3

3a 2
3g
x cosx2 sinx2
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-le
is described in Fig. 4.7
f3 = x21 sinx2 cosx2 +

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-le

68

CHAPTER 4. MULTI-BODY MECHANICAL SYSTEM

Chapter 5

Principle of Virtual work


5.1

Virtual Displacement and Virtual Work

A virtual displacement r is dened as an innitesimal and instantaneous


displacement in an arbitrary direction that does not oppose or violate constraints. Fig. 5.1 and Fig. 5.2 show examples of a particle under constrained
motion. In both gures, F(a) is the applied force, F(c) is the constraint force,
and r 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
denition of r, the virtual displacement is always tangent to the path of
motion and orthogonal to n. Hence
r n = 0
As a result, the virtual work W (c) done by a constraint force is zero, i.e.,
W (c) F(c) r = Fc n r = 0

5.2

(5.1)

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))
69

(5.2)

70

CHAPTER 5. PRINCIPLE OF VIRTUAL WORK

F(a)

n
r
Path of motion

F(c)
X

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


path

X
r

(c)

F(a)
n

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

(5.4)

Note that most of the following contents, we deal with the holonomic constraint.

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, lets 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 r of the bead is then
r =
1

dr
s js
ds

M is the number of degrees of freedom.

(5.5)

72

CHAPTER 5. PRINCIPLE OF VIRTUAL WORK

Z
s(t)

v(t)
path

Figure 5.3: Moving particle along the constrained path


where j is dened a Jacobian vector, representing a unit vector that tangent
to the path of motion, i.e.,
j=

dx
dy
dz
dr(x, y, z)
=
i+ j+ k
ds
ds
ds
ds

Specically, the virtual displacement r is related to the generalized coordinate through this Jacobian. In addition (5.5) can be expressed in a matrix
form as

T
dy
dz
s Js
(5.6)
r = dx
ds
ds
ds
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 r of the coin is
r =
=

r
q1 q1
2

i=1

r
q2 q2

r
qi
qi

r
t t

(5.8)

5.3. GENERALIZED COORDINATES AND JACOBIAN

73

q2
Z

q1

path
u(t)
r

Y
X

Figure 5.4: Moving coin on the inclined plane of the moving cart
Note that t in (5.8) is zero because of the instantaneous virtual displacement. Rewrite r in terms of spatial coordinates xyz, therefore
r =

2 

x
i=1

qi

i+

y
z
j+
k qi
qi
qi

(5.9)

Alternatively, the virtual displacement can be put in a matrix form as


r = Jq
where

(5.10)

q =

T

q1 q1

and J is the Jacobian matrix given by


x
q1
y
J = q
1
z
q1

x
q2
y
q2
z
q2

32

A system of particles
Fig. 5.5 shows a system of N -particles with K-geometric constraints. Number 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


r =

rT1 , rT2 , . . . , rTN

T

= [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

Y
X

Figure 5.5: A constrained system of particles


For a system with holonomic constraints, the position vector is also a function of generalized coordinates:
r = r (q1 , q2 , . . . , qM )
The virtual displacement r is then
r
r
r
q1 +
q2 + . . . +
qM
q1
q2
qM
M

r
=
qj
qj
j=1

r =

(5.11)

or
ri =

M

ri

qj
j=1

qj ;

i = 1, 2, . . . , N

(5.12)

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


r = Jq
where

q =

q1 q2 . . . qM

(5.13)
T
M 1

(5.14)

5.4. PRINCIPLE OF VIRTUAL WORK

75

and J is the Jacobian matrix given by

J=

J11
J21
..
.

J12
J22
..
.

...
...
..
.

J1M
J2M
..
.

JN 1 JN 2 . . . JN M

(5.15)
3N M

In (5.15), components of the Jacobian matrix Jkl =

rk
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 dierentiation 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

Lets 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)

Fi = Fi

N


(c)

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 ri = 0

i=1
2

Constraints in this case excludes the springs and dampers

(5.18)

76

CHAPTER 5. PRINCIPLE OF VIRTUAL WORK

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


forces given by
W

=
=

N


Fi ri = 0

i=1
N 


(ac)
Fi

(c)
+ Fi

(5.19)
ri = 0,

i = j

i=1

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

N

(ac)

Fi

ri = 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

DAlembert principle

Lets consider a system of N -particles and M -degrees of freedom. The Newtons second law applied to such system can be rewritten as the following
alternative form
N


(Fi miri ) = 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 ri is the acceleration of the i-th particle. The term mi ri
represents an inertia force resisting the motion of the system. From (5.21),
virtual work W done by all forces, including the inertia force mi ri , is
then zero. Or
W =

N


(Fi mi ri ) ri = 0

(5.22)

i=1

Since the virtual work done by the constraint forces is always zero, (5.22) is
simply
W  =

N 

(ac)

Fi

i=1

mi ri ri = 0

(5.23)

5.5. DALEMBERT PRINCIPLE

77

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

(ac)

Fi

M
 
ri

mi ri

i=1

k=1

qk

qk = 0

(5.24)

Rewrite(5.24) as
M


N


k=1

i=1

(ac)

(Fi

mi ri )

ri
qk

qk = 0

(5.25)

Since each qk is independent and nonzero, therefore to satisfy (5.25), each


k-term in the bracket must be zero, i.e.
N 

(ac)

Fi

 r
i

mi ri

i=1

qk

= 0;

k = 1, 2, . . . , M

(5.26)

ri
for the rest of the chapter. Equaqk
tions (5.26) are the DAlembert principle that automatically yield the equations of motion. Again the DAlembert principle can only be applied to
dynamics of the systems with holonomic constraints.
Example: Direct application of DAlembert 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 . If the unstretched length of the
springs is a, derive the equations of motion.
From the system shown in Figure 5.6, lets choose 1 , 2 , and as the
generalized coordinates. Hence
For a short notation, we dene ik

q=

1 2

T

(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 (1 , 2 , ) = 1 er

(5.28)

r2 = f (1 , 2 , ) = 2 er

(5.29)

and

78

CHAPTER 5. PRINCIPLE OF VIRTUAL WORK

m1

er

m2

2
1

Figure 5.6: Example


The accelerations of both sliders are


r1 (1 , 2 , ) = 1 2 1 er + 1 + 2 1 e
and

(5.30)

r2 (1 , 2 , ) =
2 + 2 2 er 2 + 22 e

(5.31)

The active spring forces acting on both sliders are


(ac)

F1

= k (1 a) er

(5.32)

and
(ac)

F2

= k (2 a) er

The components of the Jacobian matrix can be formulated as follows


11
12
13
21
22

r1
1
r1
2
r1

r2
1
r2
2

= er
= 0
= 1 e
= 0
= er

(5.33)

5.5. DALEMBERT PRINCIPLE


23

79

r2

= 2 e

Substitution of (5.30) to (5.33) into (5.26) yields three sets of equations:


For k = 1,


(ac)

F1
or

(ac)

m1 r1 11 + F2


m2 r2 21 = 0

(5.34)

m1 1 2 1 + K (1 a) = 0

(5.35)

For k = 2,


(ac)

F1
or

(ac)

m1 r1 12 + F2


m2 r2 22 = 0

(5.36)

m2 2 2 2 + K (2 a) = 0

(5.37)

For k = 3,


(ac)

F1
or

m1 r1 13 + F2

(ac)

m2 r2 23 = 0

m1 1 + 21 1 + m2 2 + 22 2 = 0

(5.38)

(5.39)

Equations (5.35) and (5.37) are the equations of motions. In addition, (5.39)
can be arranged as

d 
m1 21 + m2 22 = 0
(5.40)
dt
which indicates the conservation of angular momentum of the system or
o = 0.
H

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
T =

N
1
mi vi vi
2 i=1

(6.1)

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
v vdm
(6.2)
T =
2
where v is the velocity of element dm. From Figure 6.1, the velocity v of
mass dm is
(6.3)
v = vc +
Substitution of (6.3) into (6.2) yields1
T

=
=
=

The terms

1
2  (vc + ) (vc + ) dm 
1
dm + vc dm + 12
2 vc v
c
1
1
2 vc vc dm + (vc ) dm + 2

dm = 0 and
1
2

(
) ( ) dm

( ) dm

dm = m. Also

1
( ) dm =
2

Note that (A B) C = A (B C)

81

( ) dm

(6.4)

82

CHAPTER 6. LAGRANGE MECHANICS

vc
C

rc
Y

Figure 6.1: Kinetic energy of a rigid body




where ( ) 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
2
2

(6.5)

Equation (6.5) can be put in the matrix form as


1
1
T = mvcT vc + T Ic
2
2

6.2

(6.6)

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

i+
j+
k
x
y
z

(6.7)

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 =
A

F r V (rA ) V (rB )

(6.8)

Or virtual work W done by the conservative force is


W = F r = dV

(6.9)

Examples of the conservative forces F and their corresponding potential


energy V are as follows.
1. Gravity near the Earths surface
F = mgez ,

V = mgz

2. Gravitational force
F=

GM m
er ,
r2

V =

GM m
r

3. Elastic spring force


F = kx,

V =

1 2
kx
2

4. Magnetic force between two wires


F =

0 I1 I2
,
2r

V =

0 I1 I2 logr
2

where 0 = 4 107 in mks units.


5. Electric force between two charges
F =
where

Q1 Q2
,
40 r 2

V =

1
= 8.99 109 in mks units.
40

Q1 Q2
40 r

84

CHAPTER 6. LAGRANGE MECHANICS

6.3

Remarks on Properties of Generalized Coordinates for the System with Holonomic constraints

1. For a multi-degree-of-freedom dynamical system, all generalized coordinates 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 coordinates and time as
ri
ri
ri
ri
t
q1 +
q2 + . . . +
qM +
q1
q2
qM
t
M

ri
qj
=
qj
j=1

ri =

Note that t is zero because the virtual displacement is an instantari


= f (q1 , q2 , . . . , qM , t).
neous displacement, and
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 dq2
ri dqM
ri
ri dq1
+
+ ... +
+
r i =
q1 dt
q2 dt
qM dt
t
M

ri
ri
qj +
=
q
t
j
j=1
f (q1 , q2 , . . . , qM , q1 , q2 , . . . , qM , t)
where qj

dqj
is the generalized velocity.
dt

5. Remarkable observation 1
ri
r i
=
,
qk
qk

k = 1, 2, . . . , M ;

i = 1, 2, . . . , N

6.4. DERIVATION OF LAGRANGES EQUATIONS

85

6. Remarkable observation 2
d
r i
=
qk
dt

6.4

ri
qk

k = 1, 2, . . . , M ;

i = 1, 2, . . . , N

Derivation of Lagranges equations

Lets consider a system of N -particles with M -degrees of freedom. From the


DAlembert principle:
N 


(ac)

mi ri Fi

 r
i

i=1

= 0;

qk

k = 1, 2, . . . , M

(6.10)

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

i=1

miri

N

ri
ri
d
=
mi r i
qk
dt
qk
i=1

mi r i

d
dt

ri
qk



(6.11)

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


N


ri
mi ri
qk
i=1


N 

d

r i
mi r i
dt
qk

i=1



N 

d

i=1
N 


i=1

d
dt

dt qk


2




mi r i r i

Ti
d Ti

dt qk
qk

T
qk

r i
mi r i
qk




1
mi r i r i

qk 2



T
qk

where Ti is the kinetic energy of i-particle, and T is the total kinetic energy.
The second term of (6.10) is dened as generalized forces Qk . Or
Qk

N

(ac)

Fi

i=1

ri
;
qk

k = 1, 2, . . . , M

(6.12)

Consequently (6.10) becomes


d
dt

T
qk

T
= Qk ;
qk

k = 1, 2, . . . , M

(6.13)

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 dened 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
W (ac) =

N
M 

(ac)

Fi

k=1 i=1

M

ri

qk =
Qk qk
qk
k=1

(6.14)

Any generalized force Qj is therefore determined from the virtual work


done by enforcing virtual displacement in the particular j-coordinate
for one unit (qj = 1), and enforcing zero virtual displacements in the
other coordinates (qi = 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 potential energy as
(c)

Qk

N

(c)

Fi

i=1
N 


i=1

ri
qk

V
ri

(q1 , q2 , . . . , qM )
ri
qk

V
qk
(c)

Substitution of the conservative generalized forces Qk into (6.13) yields the


Lagranges equations
d
dt

T
qk

T
V
(nc)
+
= Qk ;
qk
qk

k = 1, 2, . . . , M

(6.15)

Lets dene a Lagrange function L such that


T (q, q)
V (q)
L(q, q)
An alternative form of the Lagranges equations is then
d
dt

L
qk

L
(nc)
= Qk ;
qk

k = 1, 2, . . . , M

(6.16)

6.5. EXAMPLES

87

m
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
Lagranges equation.
Solution
The pendulum shown in Fig. 6.2 has one degree of freedom. Lets choose
as the generalized coordinate. The kinetic energy T of the pendulum is
then
1
1
(6.17)
T = mv 2 = ml2 2
2
2
Also the potential energy V of the pendulum (with respect to the datum at
the hinge level) is
V = mglcos
(6.18)
There is no external forces applied to the pendulum, hence the generalized
force is zero. The Lagranges equation for the pendulum is
d
dt

V
T
+
=0

(6.19)

Substitution of (6.17) and (6.18) into (6.19) yields the equation of motion:
ml2 + mglsin = 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
e

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 = . Absolute velocity v2 of the pendulum is

x + Le
v2 = xe


= xsine

+ L e
r + xcos

(6.21)

where L is the length of the pendulum. Also




+ L
v2 v2 = x 2 sin2 + xcos

= x 2 + 2Lx cos
+ L2 2

2

(6.22)

The total kinetic energy T of the system is


1
1
T = m1 x 2 + m2 v2 v2
2
2

(6.23)

Substitute (6.22) into (6.23) to get


T =

1
1

(m1 + m2 ) x 2 + m2 Lx cos
+ m2 L2 2
2
2

(6.24)

The potential energy of the system is


V =

1 2
kx m2 gL cos
2

(6.25)

6.5. EXAMPLES

89

The Lagranges equations are


d
dt
and

T
x

V
T
+
=0
x
x

V
T
T
+
=0


Formulate each term in (6.26) and (6.27) as follows:
d
dt

d
dt

T
x

d
dt

(6.26)

(6.27)

= (m1 + m2 ) x + m2 Lcos
x

(6.28)

= (m1 + m2 ) x
+ m2 Lcos
m2 L2 sin

(6.29)

T
=0
x

(6.30)

T
= m2 Lxcos

+ m2 L2

(6.31)

= m2 L
xcos m2 Lx sin
+ m2 L2

(6.32)

= m2 Lx sin
(6.33)

V
= kx
(6.34)
x
V
= m2 gLsin
(6.35)

Plug (6.29), (6.30) and (6.32) to (6.35) into (6.26) and (6.27) to obtain the
equations of motion:

+ m2 Lcos
m2 L2 sin + kx = 0
(m1 + m2 ) x

(6.36)

xcos + m2 gLsin = 0
m2 L2 + m2 L

(6.37)

and
Lagranges equations can be used to derive equations of motion of the
rigid body or multi-body system. In this case, derivation of the Lagranges
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
1

a1

m1,
G1
I1(about G) 1
a2
l1
2 2

G2

l2
m2,
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 1 and 2 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. Lets dene 1
and 2 as the generalized coordinates. The position vector and the velocity
of G2 can be written in terms of both generalized coordinates as
rG2 = (l1 cos1 + a2 cos2 ) i + (l1 sin1 + a2 sin2 ) j
r G2 = (l1 1 sin1 a2 2 sin2 ) i + (l1 1 cos1 + a2 2 cos2 ) j

(6.38)
(6.39)

The dot product of the velocity r G2 is then


r G2 r G2 = (l1 1 sin1 a2 2 sin2 )2 + (l1 1 cos1 + a2 2 cos2 )2
= l12 21 + a22 22 + 2l1 a2 1 2 cos (1 2 )
(6.40)
Kinetic and potential energy:
The total kinetic energy T of the two-link arms is
T

= T1 + T2 

= 12 Io1 21 + 12 m2 r G2 r G2 + 12 IG2 22


!
"
= 12 I1 + m1 a21 21 + 12 m2 l12 21 + a22 22 + 2l1 a2 1 2 cos (1 2 )
+ 12 I2 22
(6.41)

6.5. EXAMPLES

91

1
1

2
2

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 Q1 and Q2 :
From Figure 6.5, the virtual work done by all nonconservative torques is
W = 1 1 2 1 + 2 2

(6.42)

Q1 = 1 2

(6.43)

Q2 = 2

(6.44)

Hence
and
Formulate the Lagranges equations as follows:
d
dt
d
dt




T
1
T
2

T
V
+
= Q1
1 1

(6.45)

T
V
+
= Q2
2 2

(6.46)

where


T
= I1 + m1 a21 1 + m2 l12 1 + m2 a2 l1 2 cos (1 2 )
1

d
dt

T
1

(6.47)

I1 + m1 a21 + m2 l12
1 + m2 a2 l1
2 cos (1 2 )

m2 a2 l1 2 ( 1 2 ) sin (1 2 )

(6.48)

92

CHAPTER 6. LAGRANGE MECHANICS


T
= m2 a2 l1 1 2 sin (1 2 )
1

(6.49)

T
= m2 a22 2 + m2 a2 l1 1 cos (1 2 ) + I2 2
2




d T
= m2 a22 + I2
2 + m2 a2 l1
1 cos (1 2 )
dt 2
m2 a2 l1 1 ( 1 2 ) sin (1 2 )

(6.50)

(6.51)

T
= m2 a2 l1 1 2 sin (1 2 )
(6.52)
2
Substitution of (6.47)-(6.52) into (6.45) and (6.46) yields the following equations of motion:


I1 + m1 a21 + m2 l12
1 + m2 a2 l1
2 cos (1 2 )
+m2 a2 l1 22 sin (1 2 ) = 1 2


(6.53)

m2 a22 + I2
2 + m2 a2 l1
1 cos (1 2 )
m2 a2 l1 21 sin (1 2 ) = 2

(6.54)

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 . Derive the dierential
equations governing the motion of the thin bar using Lagranges equations.
Solution
With the prescribed motion is constant, the number of degrees of freedom

is M = 6 N C L = 6 2 (5 + 5) 1 = 1. Lets choose as the
generalized coordinate. The angular velocity of the link 2 is
2
2 = k1 + i
2 + sinj2 + cosk2
= i

T

sin cos

(6.55)

Kinetic energy T is
1
1
T = IZ 2 + T2 I2
2
2
Substituting (6.55) into (6.56) yields
T

1
2
2 IZ

1
2

1
2
2 IZ

1
2

sin cos

2 + 2 sin2

(6.56)

I 0 0

0 I 0 sin
0 0 0
cos
(6.57)

6.5. EXAMPLES

93

k1

k2

j2

g
l

Figure 6.6: A spinning pendulum


Potential energy is V = mg(l/2)cos. The Lagranges equation can be
formulated as


V
T
d T
+
=Q
(6.58)

dt

where
T
= I

d
dt

(6.59)

= I

(6.60)

T
= I 2 sincos

(6.61)

l
V
= mg sin

(6.62)

and Q = 0. Plug (6.59)-(6.62) into (6.58) to get the equation of motion:


l
I I 2 sincos + mg sin = 0
2

(6.63)

94

CHAPTER 6. LAGRANGE MECHANICS

6.6

Lagrange Multiplier

Lagrange equation is derived from the DAlembert principle. Originally, it


can be put in the variational form as


M #

d T
k=1

dt

qk

T
V
(nc)

+
Qk
qk = 0;
qk
qk

k = 1, 2, . . . , M

(6.64)
where M is number of degrees of freedom. If all qk 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? Lets 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
k=1

dt

qk

$
p
M


T
V
(nc)

+
Qk
qk +
i
aik qk
qk
qk
i=1
k=1

=0

(6.66)
where i is called Lagrange Multiplier. Equation (6.66) can be rearranged as
M

k=1

d
dt

T
qk


T
V
(nc)
+
Qk +
i aik qk = 0
qk
qk
i=1
p

(6.67)

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 , 1 , 2 , . . . , p . If
the virtual generalized coordinates are arranged such that q1 , q2 , . . . , qM p
are independent, and qM p+1 , qM p+2 , . . . , qM are dependent. Then the
following procedure is performed to derive the equations of motion. First

6.6. LAGRANGE MULTIPLIER

95

we choose 1 , 2 , . . . , p so that each coecient in the bracket in (6.67) corresponding to qM p+1 , qM p+2 , . . . , qM is zero. Or
d
dt

T V
T
(nc) 

+
Qk + i aik = 0;
qk
qk qk
i=1
p

k = M p+1, M p+2, . . . , M

(6.68)
With the chosen i and independent q1 , q2 , . . . , qM p , the rest coecients in
the the bracket of (6.67) corresponding to q1 , q2 , . . . , qM p are all zero.
In summary, we obtain the following relation:
d
dt

T
qk


T
V
(nc)
+
= Qk
i aik ;
qk
qk
i=1
p

k = 1, 2, . . . , M

(6.69)

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 qk = 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 direction, i.e. l is not xed, this system will have two degrees of freedom. Let
r and be the generalized coordinates as shown in Figure 6.7. The kinetic
and potential energies are

1 
T = m r 2 + r 2 2
2

V = mgr cos

96

CHAPTER 6. LAGRANGE MECHANICS

Figure 6.7: A pendulum without constraint


The Lagrange equation in r-coordinate is
d
dt
d
dt

V
T
r + r = 0
(mr)
mr 2 mg cos = 0
T
r

(6.72)

The Lagrange equation in -coordinate is


d
dt
d
dt

T
+ V
= 0

mr 2 + mgr sin = 0

T


(6.73)

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


(6.72) and (6.73) together with the imposed constraint, we get



d  2 
m
r mr 2 mg cos r
mr + mgr sin + r = 0
dt

(6.74)

or


m
r mr 2 mg cos + r

d  2 
mr + mgr sin = 0
dt

(6.75)

where is the Lagrange multiplier. Choose such that

and

m
r mr 2 mg cos + = 0

(6.76)

d  2 
mr + mgr sin = 0
dt

(6.77)

6.6. LAGRANGE MULTIPLIER

97

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


solve these equations for r, and , we need another one equation which is
the constraint equation:
r=l
(6.78)
Plugging (6.78) into (6.76) and (6.77) yields
ml 2 + mg cos =

(6.79)

ml2 + mgl sin = 0

(6.80)

and
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 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. q = 0 and


= 0, where q is the vector of generalized coordinates.
q
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
qi = fi (q1 , q2 , . . . , qk , q1 , q2 , . . . , qk , t) ;

i = 1, 2, . . . , k

(7.1)

where k is the number of degrees of freedom. Lets dene the state variables
X1 , X2 , . . . , Xk , Xk+1 , Xk+2 , . . . , X2k as
X1 = q1 , X2 = q2 , . . . , Xk = qk

(7.2)

Xk+1 = q1 , Xk+2 = q2 , . . . , X2k = qk

(7.3)

and

99

100

CHAPTER 7. STABILITY ANALYSIS

Also dene 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:
= f (X1 , X2 , . . . , Xk , Xk+1 , Xk+2 , . . . , X2k , t)
X
= [f1 , f2 , . . . , f2k ]T

(7.5)

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


2 , . . . , X
k , X
k+1 , X
k+2 , . . . , X
2k , t
1 , X
(7.6)
0=f X
2 , . . . , X
2k are the set of equilibrium state determined from (7.6).
1 , X
where X
To analyze the stability, the system is initially perturbed from the equilibrium or the steady state with a small perturbation X(0) in every coordinates. Thus after the initial time, the state vector X(t) is then
+ X(t)
X(t) = X

(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



2 + X2 , . . . , X
2k + X2k , t
1 + X1 , X
+ X(t)
X
=f X

(7.9)

Equation (7.9) can be expanded using the Taylors series as




2 , . . . , X
2k , t +
1, X

+ X(t)
X
f X

f
X

X=X

X(t)

(7.10)

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


Since X
becomes


f

X(t) = AX(t)
(7.11)
X(t)

X X=X

where


f
A
X

X=X

A11
A21
..
.

A12
A22
..
.

...
...
..
.

A1,2k
A2,2k
..
.

A2k,1 A2k,2 . . . A2k,2k

(7.12)

7.2. STABILITY OF EQUILIBRIUM OR STEADY STATE

101

The component Aij in (7.12) is given by




fi
Aij =
Xj

i = 1, 2, . . . , 2k;

j = 1, 2, . . . , 2k

(7.13)

X=X

For a short notation, replace X(t) in (7.11) with Y(t). Hence the perturbation equation (7.11) becomes

Y(t)
= AY(t)

(7.14)

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

(7.15)

Substituting (7.15) into (7.14) yields


Cet = ACet

(7.16)

(A I) Cet = 0

(7.17)

Or
Thus the nontrivial solution of (7.17) is
|A I| = 0

(7.18)

From (7.18), there are 2k values of . The equilibrium or the steady state
will be unstable if either one of the following conditions is satised.
1. There exist real roots of and at least one of them is positive.
2. There exist complex roots of 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 . Determine 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 be the
generalized coordinate. Hence the absolute velocity of the bead is
v = r e
+ (rsin) ez

(7.19)

102

CHAPTER 7. STABILITY ANALYSIS

constant

smooth

er
Figure 7.1: Stability analysis of a bead steady motion
Kinetic energy T and potential energy V can be formulated as
T =


1 
1
mv v = m r 2 2 + 2 r 2 sin2
2
2

V = mgrcos

(7.20)
(7.21)

Lagrange equation is then


d
dt

V
T
+
=0

(7.22)

Formulate each term in (7.22) and substitute into the equation to get equation of motion:
m 2 r 2 sincos + mgrsin = 0
mr 2

(7.23)

Determine steady states:

Lets dene the state variables: X1 = and X2 = .


Hence the state equations are
X 1 = X2
g
X 2 = 2 sinX1 cosX1 sinX1
r

(7.24)

(7.24) can be put in the matrix form as




=
X

X 1
X 2

X2
g
2
sinX1 cosX1 sinX1
r

= f (X1 , X2 )

(7.25)

7.2. STABILITY OF EQUILIBRIUM OR STEADY STATE

103

= 0. Therefore, (7.25) becomes


For the steady state, X


0
0


2
X
1
1 cosX
1 g sinX
2 sinX
r

(7.26)

2 = 0 and
From (7.26), X


1 g
1 2 cosX
sinX
r

=0

(7.27)

1 = cos1
1 = 0 or X
There are two possible solutions for (7.27): X

g
.
2 r

Therefore the system has two steady states which are




(1)

0
0

(2)

cos1
0

g
2 r

 

(7.28)

(2) exists if and only if g 2 .


Note that the second steady state X
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


+
X(t) = X

X1 (t)
X2 (t)

(7.29)

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

Y(t)
= AY(t)
where

A11 =

f1
X1

A12

f1
=
X2


A21

f2
=
X1




X=X

X=X

X=X

A22 =

(7.30)

X2
X1

X2
=
X2




X=X

X=X

=0
=1

1
1 g cosX
= 2 cos2X
r

f2
X2

X=X

=0

104

CHAPTER 7. STABILITY ANALYSIS

Or the matrix A is


A=

0
1
1 g cosX
1 0
2 cos2X
r

(7.31)

Let the solution of (7.30) be Y(t) = Cet , can be obtained from the
characteristic equation: |A I| = 0, or
%
%

%
% 2
1 g cosX
1
% cos2X
r

%
%
%
%
%

(7.32)

Equation (7.32) yields




1,2

1
1 g cosX
= cos2X
r

1

(7.33)

1 = 0 and X
2 = 0 into (7.33) to get
For the 1st steady state, substitute X


1,2

g
=
r
2

1
2

(7.34)

g
then one of will be a positive real resulting in unstable
r
g
steady state. On the other hand if 2 then both will be conjugate
r
imaginary resulting in stable steady state. 

g
1
nd

2 = 0. Equation
and X
For the 2 steady state, X1 = cos
2r
(7.33) then becomes

In (7.34), if 2 >

"1

1 g cosX
1 2
1,2 = 2 cos2X
r
 2

1
1 2
1 sin2 X
1 g cosX
= cos2 X
r
=

g 2 4 r 2
2 r2

1
2

g
r

(7.35)

g
Since the 2nd steady state exists if and only if 2 , both 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

Anda mungkin juga menyukai