SlideShare a Scribd company logo
1 of 84
Download to read offline
Rigid Body Dynamics
9
CHAPTER OUTLINE
9.1 Introduction ..................................................................................................................................459
9.2 Kinematics....................................................................................................................................460
9.3 Equations of translational motion ...................................................................................................469
9.4 Equations of rotational motion........................................................................................................470
9.5 Moments of inertia ........................................................................................................................475
Parallel axis theorem ...........................................................................................................490
9.6 Euler’s equations...........................................................................................................................496
9.7 Kinetic energy...............................................................................................................................502
9.8 The spinning top ...........................................................................................................................504
9.9 Euler angles..................................................................................................................................510
9.10 Yaw, pitch, and roll angles ..........................................................................................................520
9.11 Quaternions ................................................................................................................................523
Problems.............................................................................................................................................532
Section 9.2 .........................................................................................................................................532
Section 9.5 .........................................................................................................................................535
Section 9.7 .........................................................................................................................................540
Section 9.8 .........................................................................................................................................540
Section 9.9 .........................................................................................................................................542
9.1 Introduction
Just as Chapter 1 provides a foundation for the dof the equations of orbital mechanics, this chapter
serves as a basis for developing the equations of satellite attitude dynamics. Chapter 1 deals with
particles, whereas here we are concerned with rigid bodies. Those familiar with rigid body dynamics
can move on to the next chapter, perhaps returning from time to time to review concepts.
The kinematics of rigid bodies is presented first. The subject depends on a theorem of the French
mathematician Michel Chasles (1793–1880). Chasles’ theorem states that the motion of a rigid body
can be described by the displacement of any point of the body (the base point) plus a rotation about a
unique axis through that point. The magnitude of the rotation does not depend on the base point. Thus,
at any instant, a rigid body in a general state of motion has an angular velocity vector whose direction
is that of the instantaneous axis of rotation. Describing the rotational component of the motion a rigid
body in three dimensions requires taking advantage of the vector nature of angular velocity and
knowing how to take the time derivative of moving vectors, which is explained in Chapter 1. Several
examples illustrate how this is done.
CHAPTER
Orbital Mechanics for Engineering Students. http://dx.doi.org/10.1016/B978-0-08-097747-8.00009-8
Copyright Ó 2014 Elsevier Ltd. All rights reserved.
459
We then move on to study the interaction between the motion of a rigid body and the forces acting on
it. Describing the translational component of the motion requires simply concentrating all of the mass at a
point, thecenter of mass, andapplying themethodsof particle mechanics to determineits motion.Indeed,
our study of the two-body problem up to this point has focused on the motion of their centers of mass
without regard to the rotational aspect. Analyzing the rotational dynamics requires computing the body’s
angular momentum, and that in turn requires accounting for how the mass is distributed throughout the
body. The mass distribution is described by the six components of the moment of inertia tensor.
Writing the equations of rotational motion relative to coordinate axes embedded in the rigid body
and aligned with the principal axes of inertia yields the nonlinear Euler equations of motion, which are
applied to a study of the dynamics of a spinning top (or one-axis gyro).
The expression for the kinetic energy of a rigid body is derived because it will be needed in the
following chapter.
The chapter next describes the two sets of three angles commonly employed to specify the orien-
tation of a body in three-dimensional space. One of these comprises the Euler angles, which are the same
as the right ascension of the node (U), argument of periapsis (u), and inclination (i) introduced in
Chapter 4 to orient orbits in space. The other set comprises the yaw, pitch, and roll angles, which are
suitable for describing the orientation of an airplane. Both the Euler angles and yaw, pitch, and roll
angles will be employed in Chapter 10.
The chapter concludes with a brief discussion of quaternions and an example of how they are used
to describe the evolution of the attitude of a rigid body.
9.2 Kinematics
Figure 9.1 shows a moving rigid body and its instantaneous axis of rotation, which defines the direction
of the absolute angular velocity vector u. The XYZ axes are a fixed, inertial frame of reference. The
position vectors RA and RB of two points on the rigid body are measured in the inertial frame. The
vector RB/A drawn from point A to point B is the position vector of B relative to A. Since the body is
rigid, RB/A has a constant magnitude even though its direction is continuously changing. Clearly,
RB ¼ RA þ RB=A
Differentiating this equation through with respect to time, we get
_
RB ¼ _
RA þ
dRB=A
dt
(9.1)
_
RA and _
RB are the absolute velocities vA and vB of points A and B. Because the magnitude of RB/A does
not change, its time derivative is given by Eqn (1.61), that is,
dRB=A
dt
¼ u  RB=A
Thus, Eqn (9.1) becomes
vB ¼ vA þ u  RB=A (9.2)
Taking the time derivative of Eqn (9.1) yields
€
RB ¼ €
RA þ
d2
RB=A
dt2
(9.3)
460 CHAPTER 9 Rigid Body Dynamics
€
RA and €
RB are the absolute accelerations aA and aB of the two points of the rigid body, while from Eqn
(1.62) we have
d2
RB=A
dt2
¼ a  RB=A þ u 

u  RB=A

in which a is the angular acceleration, a ¼ du=dt. Therefore, Eqn (9.3) can be written
aB ¼ aA þ a  RB=A þ u 

u  RB=A

(9.4)
Equations (9.2) and (9.4) are the relative velocity and acceleration formulas. Note that all quantities in
these expressions are measured in the same inertial frame of reference.
When the rigid body under consideration is connected to and moving relative to another rigid body,
computation of its inertial angular velocity u and angular acceleration a must be done with care. The
key is to remember that angular velocity is a vector. It may be found as the vector sum of a sequence of
angular velocities, each measured relative to another, starting with one measured relative to an ab-
solute frame, as illustrated in Figure 9.2. In that case, the absolute angular velocity u of body 4 is
u ¼ u1 þ u2=1 þ u3=2 þ u4=3 (9.5)
Each of these angular velocities is resolved into components along the axes of the moving frame of
reference xyz as shown in Figure 9.2, so that the vector sum may be written as
u ¼ ux
^
i þ uy
^
j þ uz
^
k (9.6)
The moving frame is chosen for convenience of the analysis, and its own inertial angular velocity is
denoted as U, as discussed in Section 1.6. According to Eqn (1.65), the absolute angular acceleration
a ¼ du=dt is obtained from Eqn (9.6) by means of the following calculation:
a ¼
du
dt

rel
þ U  u (9.7)
X
Y
Z
A
RB
RA
B
RB/A
FIGURE 9.1
A rigid body and its instantaneous axis of rotation.
9.2 Kinematics 461
where
du
dt

rel
¼ _
ux
^
i þ _
uy
^
j þ _
uz
^
k (9.8)
and _
ux ¼ dux=dt, etc.
Being able to express the absolute angular velocity vector in an appropriately chosen moving
reference frame, as in Eqn (9.6), is crucial to the analysis of rigid body motion. Once we have the
components of u, we simply differentiate them with respect to time to arrive at Eqn (9.8). Observe that
the absolute angular acceleration a and du=dtÞrel, the angular acceleration relative to the moving
frame, are the same if and only if U ¼ u. That occurs if the moving reference is a body-fixed frame,
that is, a set of xyz axes imbedded in the rigid body itself.
EXAMPLE 9.1
The airplane in Figure 9.3 flies at a constant speed v while simultaneously undergoing a constant yaw rate uyaw
about a vertical axis and describing a circular loop in the vertical plane with a radius r. The constant propeller spin
rate is uspin relative to the airframe. Find the velocity and acceleration of the tip P of the propeller relative to the
hub H, when P is directly above H. The propeller radius is l.
Solution
The xyz axes are rigidly attached to the airplane. The x-axis is aligned with the propeller’s spin axis. The y axis is
vertical, and the z axis is in the spanwise direction, so that xyz forms a right-handed triad. Although the xyz frame is
not inertial, we can imagine it to instantaneously coincide with an inertial frame.
The absolute angular velocity of the airplane has two components, the yaw rate and the counterclockwise pitch
angular velocity v/r of its rotation in the circular loop,
uairplane ¼ uyaw
^
j þ upitch
^
k ¼ uyaw
^
j þ
v
r
^
k
FIGURE 9.2
Angular velocity is the vector sum of the relative angular velocities starting with u1, measured relative to the
inertial frame.
462 CHAPTER 9 Rigid Body Dynamics
The angular velocity of the body-fixed moving frame is that of the airplane, U ¼ uairplane, so that
U ¼ uyaw
^
j þ
v
r
^
k (a)
The absolute angular velocity of the propeller is that of the airplane plus the angular velocity of the propeller
relative to the airplane
uprop ¼ uairplane þ uspin
^
i ¼ U þ uspin
^
i
which means
uprop ¼ uspin
^
i þ uyaw
^
j þ
v
r
^
k (b)
From Eqn (9.2), the velocity of point P on the propeller relative to H on the hub, vP/H, is given by
vP=H ¼ vP  vH ¼ uprop  rP=H
where rP/H is the position vector of P relative to H at this instant,
rP=H ¼ l^
j (c)
Thus, using Eqns (b) and (c),
vP=H ¼

uspin
^
i þ uyaw
^
j þ
v
r
^
k



l^
j

from which
vP=H ¼ 
v
r
l^
i þ uspinl^
k
The absolute angular acceleration of the propeller is found by substituting Eqns (a) and (b) into Eqn (9.7),
aprop ¼
duprop
dt

rel
þ U  uprop
¼

duspin
dt
^
i þ
duyaw
dt
^
j þ
dðv=rÞ
dt
^
k

þ

uyaw
^
j þ
v
r
^
k



uspin
^
i þ uyaw
^
j þ
v
r
^
k

Since uspin, uyaw, v, and r are all constant, this reduces to
aprop ¼

uyaw
^
j þ
v
r
^
k



uspin
^
i þ uyaw
^
j þ
v
r
^
k

FIGURE 9.3
Airplane with an attached xyz body frame.
9.2 Kinematics 463
Carrying out the cross product yields
aprop ¼
v
r
uspin
^
j  uyawuspin
^
k (d)
From Eqn (9.4), the acceleration of P relative to H, aP/H, is given by
aP=H ¼ aP  aH ¼ aprop  rP=H þ uprop 

uprop  rP=H

Substituting Eqns (b), (c) and (d) into this expression yields
aP=H ¼

v
r
uspin
^
j  uyawuspin
^
k



l^
j

þ

uspin
^
i þ uyaw
^
j þ
v
r
^
k



uspin
^
i þ uyaw
^
j þ
v
r
^
k



l^
j

From this we find that
aP=H ¼

uyawuspinl^
i

þ

uspin
^
i þ uyaw
^
j þ
v
r
^
k




v
r
l^
i þ uspinl^
k
¼

uyawuspinl^
i

þ

uyawuspinl^
i 

v2
r2
þ u2
spin

l^
j þ uyaw
v
r
l^
k
so that finally,
aP=H ¼ 2uyawuspinl^
i 

v2
r2
þ u2
spin

l^
j þ uyaw
v
r
l^
k
EXAMPLE 9.2
The satellite in Figure 9.4 is rotating about the z axis at a constant rate N. The xyz axes are attached to the
spacecraft, and the z-axis has a fixed orientation in inertial space. The solar panels rotate at a constant rate _
q in the
direction shown. Relative to point O, which lies at the center of the spacecraft and on the centerline of the panels,
calculate for point A on the panel
(a) Its absolute velocity and
(b) Its absolute acceleration.
FIGURE 9.4
Rotating solar panel on a rotating satellite.
464 CHAPTER 9 Rigid Body Dynamics
Solution
(a) Since the moving xyz frame is attached to the body of the spacecraft, its angular velocity is
U ¼ N^
k (a)
The absolute angular velocity of the panel is the absolute angular velocity of the spacecraft plus the angular
velocity of the panel relative to the spacecraft,
upanel ¼ _
q^
j þ N^
k (b)
The position vector of A relative to O is
rA=O ¼ 
w
2
sin q^
i þ d^
j þ
w
2
cos q^
k (c)
According to Eqn (9.2), the velocity of A relative to O is
vA=O ¼ vA  vO ¼ upanel  rA=O ¼
^
i ^
j ^
k
0 _
q N

w
2
sin q d
w
2
cos q
from which we get
vA=O ¼ 
w
2
_
q cos q þ Nd

^
i 
w
2
N sin q^
j 
w
2
_
q sin q^
k
(b) The absolute angular acceleration of the panel is found by substituting Eqns (a) and (b) into Eqn (9.7),
apanel ¼
dupanel
dt

rel
þ U  upanel
¼

d

_
q

dt
^
j þ
dN
dt
^
k
#
þ

N^
k



 _
q^
j þ N^
k

Since N and _
q are constants, this reduces to
apanel ¼ _
qN^
i (d)
To find the acceleration of A relative to O, we substitute Eqns (b)e(d) into Eqn (9.4),
aA=O ¼ aA  aO ¼ apanel  rA=O þ upanel 

upanel  rA=O

¼
^
i ^
j ^
k
_
qN 0 0

w
2
sin q d
w
2
cos q
þ

 _
q^
j þ N^
k


^
i ^
j ^
k
0 _
q N

w
2
sin q d
w
2
cos q
¼


w
2
N _
q cos q^
j þ N _
qd^
k

þ
^
i ^
j ^
k
0 _
q N

w
2
_
q cos q  Nd N
w
2
sin q 
w
2
_
q sin q
9.2 Kinematics 465
which leads to
aA=O ¼
w
2

N2
þ _
q
2

sin q^
i  N

Nd þ w _
q cos q
 ^
j 
w
2
_
q
2
cos q^
k
EXAMPLE 9.3
The gyro rotor illustrated in Figure 9.5 has a constant spin rate uspin around axis bea in the direction shown. The
XYZ axes are fixed. The xyz axes are attached to the gimbal ring, whose angle q with the vertical is increasing at a
constant rate _
q in the direction shown. The assembly is forced to precess at a constant rate N around the vertical,
For the rotor in the position shown, calculate
(a) The absolute angular velocity and
(b) The absolute angular acceleration.
Express the results in both the fixed XYZ frame and the moving xyz frame.
Solution
(a) We will need the instantaneous relationship between the unit vectors of the inertial XYZ axes and the comoving
xyz frame, which on inspecting Figure 9.6 can be seen to be
^
I ¼ cos q^
j þ sin q^
k
^
J ¼ ^
i
^
K ¼ sin q^
j þ cos q^
k
(a)
so that the matrix of the transformation from xyz to XYZ is (Section 4.5)
½QxX ¼
2
6
6
4
0 cos q sin q
1 0 0
0 sin q cos q
3
7
7
5 (b)
The absolute angular velocity of the gimbal ring is that of the base plus the angular velocity of the gimbal
relative to the base given as
ugimbal ¼ N^
K þ _
q^
i ¼ N

sin q^
j þ cos q^
k

þ _
q^
i ¼ _
q^
i þ N sin q^
j þ N cos q^
k (c)
where we made use of Eqn (a) above. Since the moving xyz frame is attached to the gimbal, U ¼ ugimbal,
so that
U ¼ _
q^
i þ N sin q^
j þ N cos q^
k (d)
The absolute angular velocity of the rotor is its spin relative to the gimbal, plus the angular velocity of the
gimbal,
urotor ¼ ugimbal þ uspin
^
k (e)
From Eqn (c), it follows that
urotor ¼ _
q^
i þ N sin q^
j þ

N cos q þ uspin

^
k (f)
466 CHAPTER 9 Rigid Body Dynamics
θ
θ
Z
FIGURE 9.6
Orientation of the fixed XZ axes relative to the rotating xz axes.
FIGURE 9.5
Rotating, precessing, nutating gyro.
9.2 Kinematics 467
Because^
i, ^
j and ^
k move with the gimbal, expression (f) is valid for any time, not just the instant shown in
Figure 9.5. Alternatively, applying the vector transformation
furotorgXYZ ¼ ½QxX furotorgxyz (g)
we obtain the angular velocity of the rotor in the inertial frame, but only at the instant shown in the figure, that
is, when the x-axis aligns with the Y-axis.
8

:
uX
uY
uZ
9



=



;
¼
2
6
6
6
4
0 cos q sin q
1 0 0
0 sin q cos q
3
7
7
7
5
8







:
_
q
N sin q
N cos q þ uspin
9



=



;
¼
8







:
N sin q cos q þ N sin q cos q þ uspin sin q
_
q
N sin2
q þ N cos2q þ uspin cos q
9



=



;
or
urotor ¼ uspin sin q^
I þ _
q^
J þ

N þ uspin cos q

^
K (h)
(b) The angular acceleration of the rotor can be found by substituting Eqns (d) and (f) into Eqn (9.7), recalling that
N, _
q, and uspin are independent of time:
arotor ¼
durotor
dt

rel
þ U  urotor
¼

d

_
q

dt
^
i þ
dðN sin qÞ
dt
^
j þ
d

N cos q þ uspin

dt
^
k
#
þ
^
i ^
j ^
k
_
q N sin q N cos q
_
q N sin q N cos q þ uspin
¼

N _
q cos q^
j  N _
q sin q^
k

þ
h
^
i

Nuspin sin q

^
j

uspin
_
q

þ ^
kð0Þ
i
Upon collecting terms, we get
arotor ¼ Nuspin sin q^
i þ _
q

N cos q  uspin
^
j  N _
q sin q^
k (i)
This expression, like Eqn (f), is valid at any time.
The components of arotor along the XYZ axes are found in the same way as for urotor,
farotorgXYZ ¼ ½QxX farotorgxyz
which means
8

:
aX
aY
aZ
9



=



;
¼
2
6
6
6
4
0 cos q sin q
1 0 0
0 sin q cos q
3
7
7
7
5
8







:
Nuspin sin q
_
q

N cos q  uspin

N _
q sin q
9




=




;
¼
8

:
N _
qcos2q þ _
quspin cos q  N _
q sin2
q
Nuspin sin q
N _
q sin q cos q  _
quspin sin q  N _
q sin q cos q
9




=




;
468 CHAPTER 9 Rigid Body Dynamics
or
arotor ¼ _
q

uspin cos q  N

^
I þ Nuspin sin q^
J  _
quspin sin q^
K (j)
Note carefully that Eqn (j) is not simply the time derivative of Eqn (h). Equations (h) and (j) are valid only at the
instant that the xyz and XYZ axes have the alignments shown in Figure 9.5.
9.3 Equations of translational motion
Figure 9.7 again shows an arbitrary, continuous, three-dimensional body of mass m. “Continuous”
means that as we zoom in on a point it remains surrounded by a continuous distribution of matter
having the infinitesimal mass dm in the limit. The point never ends up in a void. In particular, we ignore
the actual atomic and molecular microstructure in favor of this continuum hypothesis, as it is called.
Molecular microstructure does bear upon the overall dynamics of a finite body. We will use G to denote
the center of mass. The position vectors of points relative to the origin of the inertial frame will be
designated by capital letters. Thus, the position of the center of mass is RG defined as
mRG ¼
Z
m
Rdm (9.9)
R is the position of a mass element dm within the continuum. Each element of mass is acted upon by a
net external force dFnet and a net internal force dfnet. The external force comes from direct contact with
other objects and from action at a distance, such as gravitational attraction. The internal forces are
those exerted from within the body by neighboring particles. These are the forces that hold the body
together. For each mass element, Newton’s second law, Eqn (1.47), is written as
dFnet þ dfnet ¼ dm€
R (9.10)
Writing this equation for the infinite number of mass elements of which the body is composed, and
then summing them all together leads, to the integral
Z
dFnet þ
Z
dfnet ¼
Z
m
€
Rdm
FIGURE 9.7
Forces on the mass element dm of a continuous medium.
9.3 Equations of translational motion 469
Because the internal forces occur in action–reaction pairs,
R
dfnet ¼ 0. (External forces on the body are
those without an internal reactant; the reactant lies outside the body and, hence, is outside our pur-
view.) Thus,
Fnet ¼
Z
m
€
Rdm (9.11)
where Fnet is the resultant external force on the body, Fnet ¼
R
dFnet. From Eqn (9.9),
Z
m
€
Rdm ¼ m€
RG
where €
RG ¼ aG, the absolute acceleration of the center of mass. Therefore, Eqn (9.11) can be
written as
Fnet ¼ m€
RG (9.12)
We are therefore reminded that the motion of the center of mass of a body is determined solely by the
resultant of the external forces acting on it. So far, our study of orbiting bodies has focused exclusively
on the motion of their centers of mass. In this chapter, we will turn our attention to rotational motion
around the center of mass. To simplify things, we will ultimately assume that the body is not only
continuous but that it is also rigid. This means all points of the body remain at a fixed distance from
each other and there is no flexing, bending, or twisting deformation.
9.4 Equations of rotational motion
Our development of the rotational dynamics equations does not require at the outset that the body
under consideration be rigid. It may be a solid, fluid, or gas.
Point P in Figure 9.8 is arbitrary; it need not be fixed in space nor attached to a point on the body.
Then, the moment about P of the forces on mass element dm (cf. Figure 9.7) is
dMP ¼ r  dFnet þ r  dfnet
where r is the position vector of the mass element dm relative to the point P. Writing the right-hand side
as r  ðdFnet þ dfnetÞ, substituting Eqn (9.10), and integrating over all the mass elements of the body
yields
MPnet
¼
Z
m
r  €
Rdm (9.13)
where €
R is the absolute acceleration of dm relative to the inertial frame and
MPnet
¼
Z
r  dFnet þ
Z
r  dfnet
470 CHAPTER 9 Rigid Body Dynamics
But
R
r  dfnet ¼ 0 because the internal forces occur in action–reaction pairs. Thus,
MPnet
¼
Z
r  dFnet
which means the net moment includes only the moment of all of the external forces on the body.
From the product rule of calculus, we know that dðr  _
RÞ=dt ¼ r  €
R þ _
r  _
R, so that the inte-
grand in Eqn (9.13) may be written as
r  €
R ¼
d
dt

r  _
R

 _
r  _
R (9.14)
Furthermore, Figure 9.8 shows that r ¼ R  RP, where RP is the absolute position vector of P. It
follows that
_
r  _
R ¼

_
R  _
RP

 _
R ¼  _
RP  _
R (9.15)
Substituting Eqn (9.15) into Eqn (9.14), then moving that result into Eqn (9.13), yields
MPnet
¼
d
dt
Z
m
r  _
Rdm þ _
RP 
Z
m
_
Rdm (9.16)
Now, r  _
Rdm is the moment of the absolute linear momentum of mass element dm about P.
The moment of momentum, or angular momentum, of the entire body is the integral of this cross-
product over all of its mass elements. That is, the absolute angular momentum of the body relative to
point P is
HP ¼
Z
m
r  _
Rdm (9.17)
FIGURE 9.8
Position vectors of a mass element in a continuum from several key reference points.
9.4 Equations of rotational motion 471
Observing from Figure 9.8 that r ¼ rG/P þ r, we can write Eqn (9.17) as
HP ¼
Z
m

rG=P þ r

 _
Rdm ¼ rG=P 
Z
m
_
Rdm þ
Z
m
r  _
Rdm (9.18)
The last term is the absolute angular momentum relative to the center of mass G,
HG ¼
Z
r  _
Rdm (9.19)
Furthermore, by the definition of center of mass, Eqn (9.9),
Z
m
_
Rdm ¼ m _
RG (9.20)
Equations (9.19) and (9.20) allow us to write Eqn (9.18) as
HP ¼ HG þ rG=P  mvG (9.21)
This useful relationship shows how to obtain the absolute angular momentum about any point P once
HG is known.
For calculating the angular momentum about the center of mass, Eqn (9.19) can be cast in a much
more useful form by making the substitution (cf. Figure 9.8) R ¼ RG þ r, so that
HG ¼
Z
m
r 

_
RG þ _
r

dm ¼
Z
m
r  _
RGdm þ
Z
m
r  _
rdm
In the two integrals on the right, the variable is r. _
RG is fixed and can therefore be factored out of the
first integral to obtain
HG ¼
0
@
Z
m
rdm
1
A  _
RG þ
Z
m
r  _
rdm
By definition of the center of mass,
R
m
rdm ¼ 0 (the position vector of the center of mass relative to
itself is zero), which means
HG ¼
Z
m
r  _
rdm (9.22)
Since r and _
r are the position and velocity relative to the center of mass G,
R
m
r  _
rdm is the total
moment about the center of mass of the linear momentum relative to the center of mass, HGrel
. In other
words,
HG ¼ HGrel
(9.23)
This is a rather surprising fact, hidden in Eqn (9.19), and is true in general for no other point of the body.
472 CHAPTER 9 Rigid Body Dynamics
Another useful angular momentum formula, similar to Eqn (9.21), may be found by substituting
R ¼ RP þ r into Eqn (9.17),
HP ¼
Z
m
r 

_
RP þ _
r

dm ¼
0
@
Z
m
rdm
1
A  _
RP þ
Z
m
r  _
rdm (9.24)
The term on the far right is the net moment of relative linear momentum about P,
HPrel
¼
Z
m
r  _
rdm (9.25)
Also,
R
m
rdm ¼ mrG=P, where rG/P is the position of the center of mass relative to P. Thus, Eqn (9.24)
can be written as
HP ¼ HPrel
þ rG=P  mvP (9.26)
Finally, substituting this into Eqn (9.21), solving for HPrel
, and noting that vG  vP ¼ vG/P yields
HPrel
¼ HG þ rG=P  mvG=P (9.27)
This expression is useful when the absolute velocity vG of the center of mass, which is required in Eqn
(9.21), is not available.
So far, we have written down some formulas for calculating the angular momentum about an
arbitrary point in space and about the center of mass of the body itself. Let us now return to the
problem of relating angular momentum to the applied torque. Substituting Eqns (9.17) and (9.20) into
Eqn (9.16), we obtain
MPnet
¼ _
HP þ _
RP  m _
RG
Thus, for an arbitrary point P,
MPnet
¼ _
HP þ vP  mvG (9.28)
where vP and vG are the absolute velocities of points P and G, respectively. This expression is
applicable to two important special cases.
If the point P is at rest in inertial space ðvP ¼ 0Þ, then Eqn (9.28) reduces to
MPnet
¼ _
HP (9.29)
This equation holds as well if vP and vG are parallel. If P is the point of contact of a wheel rolling while
slipping in the plane. Note that the validity of Eqn (9.29) depends neither on the body’s being rigid nor
on its being in pure rotation about P. If point P is chosen to be the center of mass, then, since
vG  vG ¼ 0, Eqn (9.28) becomes
MGnet
¼ _
HG (9.30)
This equation is valid for any state of motion.
9.4 Equations of rotational motion 473
If Eqn (9.30) is integrated over a time interval, then we obtain the angular impulse–momentum
principle,
Zt2
t1
MGnet
dt ¼ HG2
 HG1
(9.31)
A similar expression follows from Eqn (9.29).
R
Mdt is the angular impulse. If the net angular impulse
is zero, then DH ¼ 0, which is a statement of the conservation of angular momentum. Keep in mind
that the angular impulse–momentum principle is not valid for just any reference point.
Additional versions of Eqns (9.29) and (9.30) can be obtained which may prove useful in special
circumstances. For example, substituting the expression for HP (Eqn (9.21)) into Eqn (9.28) yields
MPnet
¼

_
HG þ
d
dt

rG=P  mvG

þ vP  mvG
¼ _
HG þ
d
dt
½ðrG  rPÞ  mvG þ vP  mvG
¼ _
HG þ ðvG  vPÞ  mvG þ rG=P  maG þ vP  mvG
or, finally,
MPnet
¼ _
HG þ rG=P  maG (9.32)
This expression is useful when it is convenient to compute the net moment about a point other than the
center of mass. Alternatively, by simply differentiating Eqn (9.27) we get
_
HPrel
¼ _
HG þ vG=P  mvG=P
zfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflffl{
¼0
þ rG=P  maG=P
Solving for _
HG, invoking Eqn (9.30), and using the fact that aP/G ¼ aG/P leads to
MGnet
¼ _
HPrel
þ rG=P  maP=G (9.33)
Finally, if the body is rigid, the magnitude of the position vector r of any point relative to the center
of mass does not change with time. Therefore, Eqn (1.52) requires that _
r ¼ u  r, leading us to
conclude from Eqn (9.22) that
HG ¼
Z
m
r  ðu  rÞdm

Rigid body

(9.34)
Again, the absolute angular momentum about the center of mass depends only on the absolute angular
velocity and not on the absolute translational velocity of any point of the body.
474 CHAPTER 9 Rigid Body Dynamics
No such simplification of Eqn (9.17) exists for an arbitrary reference point P. However, if the
point P is fixed in inertial space and the rigid body is rotating about P, then the position vector r from
P to any point of the body is constant. It follows from Eqn (1.52) that _
r ¼ u  r. According to
Figure 9.8,
R ¼ RP þ r
Differentiating with respect to time gives
_
R ¼ _
RP þ _
r ¼ 0 þ u  r ¼ u  r
Substituting this into Eqn (9.17) yields the formula for angular momentum in this special case as
HP ¼
Z
m
r  ðu  rÞdm ðRigid body rotating about fixed point PÞ (9.35)
Although Eqns (9.34) and (9.35) are mathematically identical, one must keep in mind the notation
of Figure 9.8. Eqn (9.35) applies only if the rigid body is in pure rotation about a stationary point in
inertial space, whereas Eqn (9.34) applies unconditionally to any situation.
9.5 Moments of inertia
To use Eqn (9.29) or Eqn (9.30) to solve problems, the vectors within them have to be resolved
into components. To find the components of angular momentum, we must appeal to its definition.
We will focus on the formula for angular momentum of a rigid body about its center of mass
(Eqn (9.34)) because the expression for fixed-point rotation (Eqn (9.35)) is mathematically the
same. The integrand of Eqn (9.34) can be rewritten using the bac–cab vector identity presented in
Eqn (2.33),
r  ðu  rÞ ¼ ur2
 rðu$rÞ (9.36)
Let the origin of a comoving xyz coordinate system be attached to the center of mass G, as shown in
Figure 9.9. The unit vectors of this frame are ^
i, ^
j, and ^
k. The vectors r and u can be resolved into
components in the xyz directions to get r ¼ x^
i þ y^
j þ z^
k and u ¼ ux
^
i þ uy
^
j þ uz
^
k. Substituting these
vector expressions into the right side of Eqn (9.36) yields
r  ðu  rÞ ¼

ux
^
i þ uy
^
j þ uz
^
k

x2
þ y2
þ z2



x^
i þ y^
j þ z^
k

uxx þ uyy þ uzz

Expanding the right side and collecting the terms having the unit vectors ^
i, ^
j, and ^
k in common,
we get
r  ðu  rÞ ¼

y2
þ z2

ux  xyuy  xzuz
^
i þ  yxux þ

x2
þ z2

uy  yzuz
^
j
þ  zxux  zyuy þ

x2
þ y2

uz
^
k (9.37)
9.5 Moments of inertia 475
We put this result into the integrand of Eqn (9.34) to obtain
HG ¼ Hx
^
i þ Hy
^
j þ Hz
^
k (9.38)
where
8

:
Hx
Hy
Hz
9



=



;
¼
2
6
6
6
4
Ix Ixy Ixz
Iyx Iy Iyz
Izx Izy Iz
3
7
7
7
5
8







:
ux
uy
uz
9



=



;
(9.39a)
or, in matrix notation,
fHg ¼ ½Ifug (9.39b)
The nine components of the moment of inertia matrix [I] about the center of mass are
Ix ¼
R
y2 þ z2

dm Ixy ¼ 
R
xydm Ixz ¼ 
R
xzdm
Iyx ¼ 
R
yxdm Iy ¼
R
x2 þ z2

dm Iyz ¼ 
R
yzdm
Izx ¼ 
R
zxdm Izy ¼ 
R
zydm Iz ¼
R
x2 þ y2

dm
(9.40)
Since Iyx ¼ Ixy, Izx ¼ Ixz, and Izy ¼ Iyz, it follows that [I] is a symmetric matrix ð½IT
¼ ½IÞ. Therefore,
[I] has six independent components instead of nine. Observe that, whereas the products of inertia
Ixy, Ixz, and Iyz can be positive, negative, or zero, the moments of inertia Ix, Iy, and Iz are always
positive (never zero or negative) for bodies of finite dimensions. For this reason, [I] is a symmetric
positive-definite matrix. Keep in mind that Eqns (9.38) and (9.39) are valid as well for axes attached to
a fixed point P about which the body is rotating.
The moments of inertia reflect how the mass of a rigid body is distributed. They manifest a body’s
rotational inertia, that is, its resistance to being set into rotary motion or stopped once rotation is
FIGURE 9.9
A comoving xyz frame used to compute the moments of inertia.
476 CHAPTER 9 Rigid Body Dynamics
underway. It is not an object’s mass alone but how that mass is distributed that determines how the
body will respond to applied torques.
If the xy plane is a plane of symmetry, then for any x and y within the body there are identical mass
elements located at þz and z. This means the products of inertia with z in the integrand vanish.
Similar statements are true if xz or yz are symmetry planes. In summary, we conclude
If the xy plane is a plane of symmetry of the body, then Ixz ¼ Iyz ¼ 0.
If the xz plane is a plane of symmetry of the body, then Ixy ¼ Iyz ¼ 0.
If the yz plane is a plane of symmetry of the body, then Ixy ¼ Ixz ¼ 0.
It follows that if the body has two planes of symmetry relative to the xyz frame of reference, then all
three products of inertia vanish, and [I] becomes a diagonal matrix such that,
½I ¼
2
6
6
4
A 0 0
0 B 0
0 0 C
3
7
7
5 (9.41)
A, B, and C are the principal moments of inertia (all positive), and the xyz axes are the body’s
principal axes of inertia. In this case, relative to either the center of mass or a fixed point of rotation,
we have
Hx ¼ Aux Hy ¼ Buy Hz ¼ Cuz (9.42)
In general, the angular velocity u and the angular momentum H are not parallel. However, if, for
example, u ¼ u^
i, then according to Eqn (9.42), H ¼ Au. In other words, if the angular velocity is
aligned with a principal direction, so is the angular momentum. In that case, the two vectors u and H
are indeed parallel.
Each of the three principal moments of inertia can be expressed as follows:
A ¼ mk2
x B ¼ mk2
y C ¼ mk2
z (9.43)
where m is the mass of the body and kx, ky, and kz are the three radii of gyration. One may imagine
the mass of a body to be concentrated around a principal axis at a distance equal to the radius of
gyration.
The moments of inertia for several common shapes are listed in Figure 9.10. By symmetry, their
products of inertia vanish for the coordinate axes used. Formulas for other solid geometries can be
found in engineering handbooks and in dynamics textbooks.
For a mass concentrated at a point, the moments of inertia in Eqn (9.40) are just the mass times
the integrand evaluated at the point. That is, the moment of inertia matrix [I(m)
] of a point mass m is
given by
h
IðmÞ
i
¼
2
6
6
4
m

y2 þ z2

mxy mxz
mxy m

x2 þ z2

myz
mxz myz m

x2 þ y2

3
7
7
5 (9.44)
9.5 Moments of inertia 477
EXAMPLE 9.4
The following table lists out the mass and coordinates of seven point masses. Find the center of mass of the system
and the moments of inertia about the origin.
Solution
The total mass of this system is
m ¼
X
7
i¼1
mi ¼ 28 kg
For concentrated masses, the integral in Eqn (9.9) is replaced by the mass times its position vector. Therefore, in
this case, the three components of the position vector of the center of mass are xG ¼ ð1=mÞ
P7
i¼1 mixi,
yG ¼ ð1=mÞ
P7
i¼1 miyi, and zG ¼ ð1=mÞ
P7
i¼1 mizi, so that
xG ¼ 0:35 m yG ¼ 0:01964 m zG ¼ 0:4411 m
Point, i Mass mi (kg) xi (m) yi (m) zi (m)
1 3 0.5 0.2 0.3
2 7 0.2 0.75 0.4
3 5 1 0.8 0.9
4 6 1.2 1.3 1.25
5 2 1.3 1.4 0.8
6 4 0.3 1.35 0.75
7 1 1.5 1.7 0.85
(b)
(a) (c)
FIGURE 9.10
Moments of inertia for three common homogeneous solids of mass m. (a) Solid circular cylinder. (b) Circular
cylindrical shell. (c) Rectangular parallelepiped.
478 CHAPTER 9 Rigid Body Dynamics
The total moment of inertia is the sum over all the particles of Eqn (9.44) evaluated at each point. Thus,
½I ¼
2
4
0:39 0:3 0:45
0:3 1:02 0:18
0:45 0:18 0:87
3
5
zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{
ð1Þ
þ
2
4
5:0575 1:05 0:56
1:05 1:4 2:1
0:56 2:1 4:2175
3
5
zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{
ð2Þ
þ
2
4
7:25 4 4:5
4 9:05 3:6
4:5 3:6 8:2
3
5
zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{
ð3Þ
þ
2
4
19:515 9:36 9
9:36 18:015 9:75
9 9:75 18:78
3
5
zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{
ð4Þ
þ
2
4
5:2 3:64 2:08
3:64 4:66 2:24
2:08 2:24 7:3
3
5
zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{
ð5Þ
þ
2
4
9:54 1:62 0:9
1:62 2:61 4:05
0:9 4:05 7:65
3
5
zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{
ð6Þ
þ
2
4
3:6125 2:55 1:275
2:55 2:9725 1:445
1:275 1:445 5:14
3
5
zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{
ð7Þ
or
½I ¼
2
6
6
4
50:56 20:42 14:94
20:42 39:73 14:90
14:94 14:90 52:16
3
7
7
5

kg$m2

EXAMPLE 9.5
Calculate the moments of inertia of a slender, homogeneous straight rod of length [ and mass m shown in
Figure 9.11. One end of the rod is at the origin and the other has coordinates (a, b, c).
Solution
A slender rod is one whose cross-sectional dimensions are negligible compared with its length. The mass is
concentrated along its centerline. Since the rod is homogeneous, the mass per unit length r is uniform and given by
r ¼
m
‘
(a)
The length of the rod is
‘ ¼
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
a2 þ b2 þ c2
p
(a, b, c)
(0, 0, 0)
A
B
FIGURE 9.11
A uniform slender bar of mass m and length [.
9.5 Moments of inertia 479
Starting with Ix, we have from Eqn (9.40),
Ix ¼
Z‘
0

y2
þ z2

rds
in which we replaced the element of mass dm by rds, where ds is the element of length along the rod. The distance
s is measured from end A of the rod, so that the x, y, and z coordinates of any point along it are found in terms of s by
the following relations:
x ¼
s
‘
a y ¼
s
‘
b z ¼
s
‘
c
Thus,
Ix ¼
Z‘
0

s2
‘2
b2
þ
s2
‘2
c2

rds ¼ r
b2 þ c2
‘2
Z‘
0
s2
ds ¼
1
3
r

b2
þ c2

‘
Substituting Eqn (a) yields
Ix ¼
1
3
m

b2
þ c2

In precisely the same way, we find
Iy ¼
1
3
m

a2
þ c2

Iz ¼
1
3
m

a2
þ b2

For Ixy we have
Ixy ¼ 
Z‘
0
xyrds ¼ 
Z‘
0
s
‘
a $
s
‘
brds ¼ r
ab
‘2
Z‘
0
s2
ds ¼ 
1
3
rab‘
Once again using Eqn (a),
Ixy ¼ 
1
3
mab
Likewise,
Ixz ¼ 
1
3
mac Iyz ¼ 
1
3
mbc
EXAMPLE 9.6
The gyro rotor (Figure 9.12) in Example 9.3 has a mass m of 5 kg, radius r of 0.08 m, and thickness t of 0.025 m.
If N ¼ 2.1 rad/s, _
q ¼ 4 rad=s, u ¼ 10.5 rad/s, and q ¼ 60, calculate
(a) The angular momentum of the rotor about its center of mass G in the body-fixed xyz frame.
(b) The angle between the rotor’s angular velocity vector and its angular momentum vector.
480 CHAPTER 9 Rigid Body Dynamics
Solution
Equation (f) from Example 9.3 gives the components of the absolute angular velocity of the rotor in the moving xyz
frame.
ux ¼ _
q ¼ 4 rad=s
uy ¼ N sin q ¼ 2:1$sin 60 ¼ 1:819 rad=s
uz ¼ uspin þ N cos q ¼ 10:5 þ 2:1$cos 60 ¼ 11:55 rad=s
(a)
Therefore,
u ¼ 4^
i þ 1:819^
j þ 11:55^
k ðrad=sÞ (b)
All three coordinate planes of the body-fixed xyz frame contain the center of mass G and all are planes of symmetry
of the circular cylindrical rotor. Therefore, [xy ¼ [zx ¼ [yz ¼ 0.
From Figure 9.10(a), we see that the nonzero diagonal entries in the moment of inertia tensor are
A ¼ B ¼
1
12
mt2
þ
1
4
mr2
¼
1
12
$5$0:0252
þ
1
4
$5$0:082
¼ 0:008260 kg$m2
C ¼
1
2
mr2
¼
1
2
$5$0:082
¼ 0:0160 kg$m2
(c)
We can use Eqn (9.42) to calculate the angular momentum, because the origin of the xyz frame is the rotor’s
center of mass (which in this case also happens to be a fixed point of rotation, which is another reason why we can
use Eqn (9.42)). Substituting Eqns (a) and (c) in Eqn (9.42) yields
Hx ¼ Aux ¼ 0:008260$4 ¼ 0:03304 kg$m2=s
Hy ¼ Buy ¼ 0:008260$1:819 ¼ 0:0150 kg$m2=s
Hz ¼ Cuz ¼ 0:0160$11:55 ¼ 0:1848 kg$m2=s
(d)
so that
H ¼ 0:03304^
i þ 0:0150^
j þ 0:1848^
k

kg$m2=s

(e)
θ
ω
FIGURE 9.12
Rotor of the gyroscope in Figure 9.4.
9.5 Moments of inertia 481
The angle f between H and u is found by taking the dot product of the two vectors,
f ¼ cos1

H$u
Hu

¼ cos1

2:294
0:1883$12:36

¼ 9:717 (f)
As this problem illustrates, the angular momentum and the angular velocity are in general not collinear.
Consider a coordinate system x0y0z0 with the same origin as xyz but a different orientation. Let [Q]
be the orthogonal matrix ð½Q1
¼ ½QT
Þ that transforms the components of a vector from the xyz
system to the x0y0z0 frame. Recall from Section 4.5 that the rows of [Q] are the direction cosines of
the x0y0z0 axes relative to xyz. If fH0
g comprises the components of the angular momentum vector
along the x0y0z0 axes, then fH0
g is obtained from its components {H} in the xyz frame by the
relation
fH0
g ¼ ½QfHg
From Eqn (9.39), we can write this as
fH0
g ¼ ½Q½Ifug (9.45)
where [I] is the moment of inertia matrix (Eqn (9.39)) in the xyz coordinates. Like the angular mo-
mentum vector, the components {u} of the angular velocity vector in the xyz system are related to
those in the primed system ðfu0gÞ by the expression
fu0
g ¼ ½Qfug
The inverse relation is simply
fug ¼ ½Q1
fu0
g ¼ ½QT
fu0
g (9.46)
Substituting this into Eqn (9.45), we get
fH0
g ¼ ½Q½I½QT
fu0
g (9.47)
But the components of angular momentum and angular velocity in the x0y0z0 frame are related by an
equation of the same form as Eqn (9.39), so that
fH0
g ¼ ½I0
fu0
g (9.48)
where ½I0
 comprises the components of the inertia matrix in the primed system. Comparing the right-
hand sides of Eqns (9.47) and (9.48), we conclude that
½I0
 ¼ ½Q½I½QT
(9.49a)
that is,
2
6
6
4
Ix0 Ix0y0 Ix0z0
Iy0x0 Iy0 Iy0z0
Iz0x0 Iz0y0 Iz0
3
7
7
5 ¼
2
6
6
4
Q11 Q12 Q13
Q21 Q22 Q23
Q31 Q32 Q33
3
7
7
5
2
6
6
4
Ix Ixy Ixz
Iyx Iy Iyz
Izx Izy Iz
3
7
7
5
2
6
6
4
Q11 Q21 Q31
Q12 Q22 Q32
Q13 Q32 Q33
3
7
7
5 (9.49b)
482 CHAPTER 9 Rigid Body Dynamics
This shows how to transform the components of the inertia matrix from the xyz coordinate system to
any other orthogonal system with a common origin. Thus, for example,
Ix0 ¼ P Q11 Q12 Q13 R
zfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflffl{
PRow 1R
2
6
6
4
Ix Ixy Ixz
Iyx Iy Iyz
Izx Izy Iz
3
7
7
5
8





:
Q11
Q12
Q13
9


=


;
zfflfflfflfflffl}|fflfflfflfflffl{
PRow 1R
T
Iy0z0 ¼ P Q21 Q22 Q23 R
zfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflffl{
PRow 2R
2
6
6
4
Ix Ixy Ixz
Iyx Iy Iyz
Izx Izy Izz
3
7
5
8



:
Q31
Q32
Q33
9

=

;
zfflfflfflfflffl}|fflfflfflfflffl{
PRow 3R
T
(9.50)
etc.
Any object represented by a square matrix whose components transform according to Eqn (9.49) is
called a second-order tensor. We may therefore refer to [I] as the inertia tensor.
EXAMPLE 9.7
Find the mass moment of inertia of the system of point masses in Example 9.4 about an axis from the origin
through the point with coordinates (2 m, 3 m, 4 m).
Solution
From Example 9.4, the moment of inertia tensor for the system of point masses is
½I ¼
2
6
6
4
50:56 20:42 14:94
20:42 39:73 14:90
14:94 14:90 52:16
3
7
7
5

kg$m2

The vector V connecting the origin with (2 m, 3 m, 4 m) is
V ¼ 2^
i  3^
j þ 4^
k
The unit vector in the direction of V is
^
uV ¼
V
kVk
¼ 0:3714^
i  0:5571^
j þ 0:7428^
k
We may consider ^
uV as the unit vector along the x0-axis of a rotated Cartesian coordinate system. Then, from Eqn
(9.50) we get
‘V 0 ¼ P 0:3714 0:5571 0:7428 R
2
6
6
6
6
4
50:56 20:42 14:94
20:42 39:73 14:90
14:94 14:90 52:16
3
7
7
7
7
5
8









:
0:3714
0:5571
0:7428
9




=




;
¼ P 0:3714 0:5571 0:7428 R
8









:
3:695
3:482
24:90
9




=




;
¼ 19:06 kg$m2
9.5 Moments of inertia 483
EXAMPLE 9.8
For the satellite of Example 9.2, which is reproduced in Figure 9.13, the data are as follows: N ¼ 0.1 rad/s and
_
q ¼ 0:01 rad=s, in the directions shown: q ¼ 40 and d0 ¼ 1.5 m. The length, width, and thickness of the panel
are [ ¼ 6 m, w ¼ 2 m, and t ¼ 0.025 m. The uniformly distributed mass of the panel is 50 kg. Find the angular
momentum of the panel relative to the center of mass O of the satellite.
Solution
We can treat the panel as a thin parallelepiped. The panel’s xyz axes have their origin at the center of mass G of the
panel and are parallel to its three edge directions. According to Figure 9.10(c), the moments of inertia relative to
the xyz coordinate system are
‘Gx ¼
1
12
m

‘2
þ t2

¼
1
12
$50$

62
þ 0:0252

¼ 150:0 kg$m2
‘Gy ¼
1
12
m

w2
þ t2

¼
1
12
$50$

22
þ 0:0252

¼ 16:67 kg$m2
‘Gz ¼
1
12
m

w2
þ ‘2

¼
1
12
$50$

22
þ 62

¼ 166:7 kg$m2
‘Gxy ¼ ‘Gxz ¼ ‘Gyz ¼ 0
(a)
In matrix notation,
½lG ¼
2
6
6
4
150:0 0 0
0 16:67 0
0 0 166:7
3
7
7
5

kg$m2

(b)
The unit vectors of the satellite’s x0y0z0 system are related to those of the panel’s xyz frame. By inspection,
^
i
0
¼ sin q^
i þ cos q^
k ¼ 0:6428^
i þ 0:7660^
k
^
j
0
¼ ^
j
^
k
0
¼ cos q^
i þ sin q^
k ¼ 0:7660^
i þ 0:6428^
k
(c)
θ
FIGURE 9.13
A satellite and solar panel.
484 CHAPTER 9 Rigid Body Dynamics
The matrix [Q] of the transformation from xyz to x0y0z0 comprises the direction cosines of ^
i
0
, ^
j
0
, and ^
k
0
:
½Q ¼
2
6
6
4
0:6428 0 0:7660
0 1 0
0:7660 0 0:6428
3
7
7
5 (d)
In Example 9.2, we found that the absolute angular velocity of the panel, in the satellite’s x0y0z0 frame of
reference, is
u ¼ _
q^
j
0
þ N^
k
0
¼ 0:01^
j
0
þ 0:1^
k
0
ðrad=sÞ
That is,
fu0
g ¼
8

:
0
0:01
0:1
9


=


;
ðrad=sÞ (e)
To find the absolute angular momentum fH0
Gg in the satellite system requires the use of Eqn (9.39),
fH0
Gg ¼ ½I0
Gfu0
g (f)
Before doing so, we must transform the components of the moments of the inertia tensor in Eqn (b) from the
unprimed system to the primed system, by means of Eqn (9.49),
½I0
G ¼ ½Q½IG½QT
¼
2
6
6
4
0:6428 0 0:7660
0 1 0
0:7660 0 0:6428
3
7
7
5
2
6
6
4
150:0 0 0
0 16:67 0
0 0 166:7
3
7
7
5
2
6
6
4
0:6428 0 0:7660
0 1 0
0:7660 0 0:6428
3
7
7
5
so that
½I0
G ¼
2
6
6
4
159:8 0 8:205
0 16:67 0
8:205 0 156:9
3
7
7
5

kg$m2

(g)
Then Eqn (f) yields
fH0
Gg ¼
2
6
6
4
159:8 0 8:205
0 16:67 0
8:205 0 156:9
3
7
7
5
8





:
0
0:01
0:1
9


=


;
¼
8





:
0:8205
0:1667
15:69
9


=


;

kg$m2
=s

or, in vector notation,
HG ¼ 0:8205^
i
0
 0:1667^
j
0
þ 15:69^
k
0

kg$m2
=s

(h)
This is the absolute angular momentum of the panel about its own center of mass G, and it is used in Eqn (9.27) to
calculate the angular momentum HOrel
relative to the satellite’s center of mass O,
HOrel
¼ HG þ rG=O  mvG=O (i)
rG/O is the position vector from O to G,
rG=O ¼

dO þ
‘
2

^
j
0
¼

1:5 þ
6
2

^
j
0
¼ 4:5^
j
0
ðmÞ (j)
9.5 Moments of inertia 485
The velocity of G relative to O, vG/O, is found from Eqn (9.2),
vG=O ¼ usatellite  rG=O ¼ N^
k
0
 rG=O ¼ 0:1^
k
0
 4:5^
j
0
¼ 0:45^
i
0
ðm=sÞ (k)
Substituting Eqns (h), (j) and (k) into Eqn (i) finally yields
HOrel
¼

0:8205^
i
0
 0:1667^
j
0
þ 15:69^
k
0

þ 4:5^
j
0

h
50

0:45^
i
0
i
¼ 0:8205^
i
0
 0:1667^
j
0
þ 116:9^
k
0 
kg$m2=s

(l)
Note that we were unable to use Eqn (9.21) to find the absolute angular momentum HO because that requires
knowing the absolute velocity vG, which in turn depends on the absolute velocity of O, which was not provided.
How can we find the direction cosine matrix [Q] such that Eqn (9.49) will yield a moment of inertia
matrix ½I0
 that is diagonal, that is, of the form given by Eqn (9.41)? In other words, how do we find the
principal directions (“eigenvectors”) and the corresponding principal values (“eigenvalues”) of the
moment of inertia tensor?
Let the angular velocity vector u be parallel to the principal direction defined by the vector e, so
that u ¼ be, where b is a scalar. Since u points in the principal direction of the inertia tensor, so must
H, which means H is also parallel to e. Therefore, H ¼ ae, where a is a scalar. From Eqn (9.39), it
follows that
afeg ¼ ½IðbfegÞ
or
½Ifeg ¼ lfeg
where l ¼ a/b (a scalar). That is,
2
4
Ix Ixy Ixz
Ixy Iy Iyz
Ixz Iyz Iz
3
5
8

:
ex
ey
ez
9
=
;
¼ l
8

:
ex
ey
ez
9
=
;
This can be written
2
4
Ix  l Ixy Ixz
Ixy Iy  l Iyz
Ixz Iyz Iz  l
3
5
8

:
ex
ey
ez
9
=
;
¼
8

:
0
0
0
9
=
;
(9.51)
The trivial solution of Eqn (9.51) is e ¼ 0, which is of no interest. The only way that Eqn (9.51) will
not yield the trivial solution is if the coefficient matrix on the left is singular. That will occur if its
determinant vanishes, that is, if
Ix  l Ixy Ixz
Ixy Iy  l Iyz
Ixz Iyz Iz  l
¼ 0 (9.52)
486 CHAPTER 9 Rigid Body Dynamics
Expanding the determinant, we find
Ix  l Ixy Ixz
Ixy Iy  l Iyz
Ixz Iyz Iz  l
¼ l3
þ J1l2
 J2l þ J3 (9.53)
where
J1 ¼ Ix þ Iy þ Iz
J2 ¼
Ix Ixy
Ixy Iy
þ
Ix Ixz
Ixz Iz
þ
Iy Iyz
Iyz Iz
J3 ¼
Ix Ixy Ixz
Ixy Iy Iyz
Ixz Iyz Iz
(9.54)
J1, J2, and J3 are invariants, that is, they have the same value in every Cartesian coordinate system.
Equations (9.52) and (9.53) yield the characteristic equation of the tensor [I],
l3
 J1l2
þ J2l  J3 ¼ 0 (9.55)
The three roots lp (p ¼ 1, 2, 3) of this cubic equation are real, since [I] is symmetric; furthermore, they
are all positive, since [I] is a positive-definite matrix. We substitute each root, or eigenvalue, lp back
into Eqn (9.51) to obtain
2
6
6
4
Ix  lp Ixy Ixz
Ixy Iy  lp Iyz
Ixz Iyz Iz  lp
3
7
7
5
8







:
e
ðpÞ
x
e
ðpÞ
y
e
ðpÞ
z
9



=



;
¼
8





:
0
0
0
9


=


;
; p ¼ 1; 2; 3 (9.56)
Solving this system yields the three eigenvectors e(p)
corresponding to each of the three eigenvalues lp.
The three eigenvectors are orthogonal, also due to the symmetry of the matrix [I]. Each eigenvalue is a
principal moment of inertia, and its corresponding eigenvector is a principal direction.
EXAMPLE 9.9
Find the principal moments of inertia and the principal axes of inertia of the inertia tensor.
½I ¼
2
6
6
4
100 20 100
20 300 50
100 50 500
3
7
7
5 kg$m2
9.5 Moments of inertia 487
Solution
We seek the nontrivial solutions of the system [I]{e} ¼ l{e}, that is,
2
6
6
4
100  l 20 100
20 300  l 50
100 50 500  l
3
7
7
5
8





:
ex
ey
ez
9


=


;
¼
8





:
0
0
0
9


=


;
(a)
From Eqn (9.54),
J1 ¼ 100 þ 300 þ 500 ¼ 900
J2 ¼
100 20
20 300
þ
100 100
100 500
þ
300 50
50 500
¼ 217; 100
J3 ¼
100 20 100
20 300 50
100 50 500
¼ 11; 350; 000
(b)
Thus, the characteristic equation is
l3
 900l2
þ 217;100l  11;350;000 ¼ 0 (c)
The three roots are the principal moments of inertia, which are found to be
l1 ¼ 532:052 l2 ¼ 295:840 l3 ¼ 72:1083 (d)
Each of these is substituted, in turn, back into Eqn (a) to find its corresponding principal direction.
Substituting l1 ¼ 532:052 kg$m2 into Eqn (a) we obtain
2
6
6
4
432:052 20:0000 100:0000
20:0000 232:052 50:0000
100:0000 50:0000 32:0519
3
7
7
5
8







:
e
ð1Þ
x
e
ð1Þ
y
e
ð1Þ
z
9



=



;
¼
8





:
0
0
0
9


=


;
(e)
Since the determinant of the coefficient matrix is zero, at most two of the three equations in Eqn (e) are inde-
pendent. Thus, at most, two of the three components of the vector e(1)
can be found in terms of the third. We can
therefore arbitrarily set e
ð1Þ
x ¼ 1 and solve for e
ð1Þ
y and e
ð1Þ
z using any two of the independent equations in Eqn (e).
With e
ð1Þ
x ¼ 1, the first two of Eqn (e) become
20:0000e
ð1Þ
y  100:000e
ð1Þ
z ¼ 432:052
232:052e
ð1Þ
y  50:000e
ð1Þ
z ¼ 20:0000
(f)
Solving these two equations for e
ð1Þ
y and e
ð1Þ
z yields, together with the assumption that e
ð1Þ
x ¼ 1,
e
ð1Þ
x ¼ 1:00000 e
ð1Þ
y ¼ 0:882793 e
ð1Þ
z ¼ 4:49708 (g)
The unit vector in the direction of e(1)
is
^
e1 ¼
eð1Þ

eð1Þ

 ¼
1:00000^
i þ 0:882793^
j  4:49708^
k
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1:000002 þ 0:8827932 þ ð4:49708Þ2
q
488 CHAPTER 9 Rigid Body Dynamics
or
^
e1 ¼ 0:213186^
i þ 0:188199^
j  0:958714^
k

l1 ¼ 532:052 kg$m2

(h)
Substituting l2 ¼ 295:840 kg$m2 into Eqn (a) and proceeding as above we find that
^
e2 ¼ 0:176732^
i  0:972512^
j  0:151609^
k

l2 ¼ 295:840 kg$m2

(i)
The two unit vectors ^
e1 and ^
e2 define two of the three principal directions of the inertia tensor. Observe that
^
e1$^
e2 ¼ 0, as must be the case for symmetric matrices.
To obtain the third principal direction ^
e3, we can substitute l3 ¼ 72:1083 kg$m2 into Eqn (a) and pro-
ceed as above. However, since the inertia tensor is symmetric, we know that the three principal directions
are mutually orthogonal, which means ^
e3 ¼ ^
e1  ^
e2. Substituting Eqns (h) and (i) into the crossproduct, we
find that
^
e3 ¼ 0:960894^
i  0:137114^
j  0:240587^
k

l3 ¼ 72:1083 kg$m2

(j)
We can check our work by substituting l3 and ^
e3 into Eqn (a) and verify that it is indeed satisfied:
2
6
6
4
100  72:1083 20 100
20 300  72:1083 50
100 50 500  72:1083
3
7
7
5
8





:
0:960894
0:137114
0:240587
9


=


;
¼
O
8





:
0
0
0
9


=


;
(k)
The components of the vectors ^
e1, ^
e2, and ^
e3 define the three rows of the orthogonal transformation [Q] from the
xyz system into the x0y0z0 system aligned along the three principal directions:
½Q ¼
2
6
6
4
0:213186 0:188199 0:958714
0:176732 0:972512 0:151609
0:960894 0:137114 0:240587
3
7
7
5 (l)
Indeed, if we apply the transformation in Eqn (9.49), ½I0
 ¼ ½Q½I½QT
, we find
½I0
 ¼
2
6
6
4
0:213186 0:188199 0:958714
0:176732 0:972512 0:151609
0:960894 0:137114 0:240587
3
7
7
5
2
6
6
4
100 20 100
20 300 50
100 50 500
3
7
7
5

2
6
6
4
0:213186 0:176732 0:960894
0:188199 0:972512 0:137114
0:958714 0:151609 0:240587
3
7
7
5
¼
2
6
6
4
532:052 0 0
0 295:840 0
0 0 72:1083
3
7
7
5

kg$m2

9.5 Moments of inertia 489
An alternative to the above hand calculations in Example 9.9 is to type the following lines in the
MATLAB Command Window:
I ¼ [ 100 -20 -100
-20 300 -50
-100 -50 500];
[eigenVectors, eigenValues] ¼ eig(I)
Hitting the Enter (or Return) key yields the following output to the Command Window:
eigenVectors ¼
0.9609 0.1767 -0.2132
0.1371 -0.9725 -0.1882
0.2406 -0.1516 0.9587
eigenValues ¼
72.1083 0 0
0 295.8398 0
0 0 532.0519
Two of the eigenvectors delivered by MATLAB are opposite in direction to those calculated in
Example 9.9. This illustrates the fact that we can determine an eigenvector only to within an arbitrary
scalar factor. That is, suppose e is an eigenvector of the tensor [I] so that [I]{e} ¼ l{e}. Multiplying
this equation through by an arbitrary scalar a yields [I]{e}a ¼ l{e}a, or [I]{ae} ¼ l{ae}, which means
that {ae} is an eigenvector corresponding to the same eigenvalue l.
Parallel axis theorem
Suppose the rigid body in Figure 9.14 is in pure rotation about point P. Then, according to Eqn (9.39),
fHPrel
g ¼ ½IPfug (9.57)
η
ζ
ξ
FIGURE 9.14
The moments of inertia are to be computed at P, given their values at G.
490 CHAPTER 9 Rigid Body Dynamics
where [IP] is the moment of inertia tensor about P, given by Eqn (9.40) with
x ¼ xG=P þ x y ¼ yG=P þ h z ¼ zG=P þ z
On the other hand, we have from Eqn (9.27) that
HPrel
¼ HG þ rG=P  mvG=P (9.58)
The vector rG/P  mvG/P is the angular momentum about P of the concentrated mass m located at the
center of mass G. Using matrix notation, it is computed as follows:
n
rG=P  mvG=P
o
hfH
ðmÞ
Prel
g ¼
h
I
ðmÞ
P
i
fug (9.59)
where
h
I
ðmÞ
P
i
, the moment of inertia of the point mass m about P, is obtained from Eqn (9.44), with
x ¼ xG/P, y ¼ yG/P, and z ¼ zG/P. That is,
½I
ðmÞ
P  ¼
2
6
6
6
4
m

y2
G=P þ z2
G=P

mxG=P yG=P mxG=P zG=P
mxG=P yG=P m

x2
G=P þ z2
G=P

myG=P zG=P
mxG=P zG=P myG=P zG=P m

x2
G=P þ y2
G=P

3
7
7
7
5
(9.60)
Of course, Eqn (9.39) requires
fHGg ¼ ½IGfug
Substituting this together with Eqns (9.57) and (9.59) into Eqn (9.58) yields
½IPfug ¼ ½IGfug þ
h
I
ðmÞ
P
i
fug ¼

½IG þ
h
I
ðmÞ
P
i
fug
From this, we may infer the parallel axis theorem,
½IP ¼ ½IG þ
h
I
ðmÞ
P
i
(9.61)
The moment of inertia about P is the moment of inertia about the parallel axes through the center of
mass plus the moment of inertia of the center of mass about P. That is,
IPx
¼ IGx
þ m

y2
G=P þ z2
G=P

IPy
¼ IGy
þ m

y2
G=P þ x2
G=P

IPz
¼ IGz
þ m

x2
G=P þ y2
G=P

IPxy
¼ IGxy
 mxG=P yG=P IPxz
¼ IGxz
 mxG=P zG=P IPyz
¼ IGyz
 myG=P zG=P
(9.62)
9.5 Moments of inertia 491
EXAMPLE 9.10
Find the moments of inertia of the rod in Example 9.5 (Figure 9.15) about its center of mass G.
Solution
From Example 9.5,
½IA ¼
2
6
6
6
6
6
6
6
6
6
4
1
3
m

b2
þ c2


1
3
mab 
1
3
mac

1
3
mab
1
3
m

a2
þ c2


1
3
mbc

1
3
mac 
1
3
mbc
1
3
m

a2
þ b2

3
7
7
7
7
7
7
7
7
7
5
Using Eqn (9.62)1, and noting the coordinates of the center of mass in Figure 9.15,
IGx
¼ IAx
 m
h
ðyG  0Þ2
þ ðzG  0Þ2
i
¼
1
3
m

b2
þ c2

 m

b
2
2
þ
c
2
2
¼
1
12
m

b2
þ c2

Equation (9.62)4 yields
IGxy
¼ IAxy
þ mðxG  0ÞðyG  0Þ ¼ 
1
3
mab þ m$
a
2
$
b
2
¼ 
1
12
mab
The remaining four moments of inertia are found in a similar fashion, so that
½IG ¼
2
6
6
6
6
6
6
6
6
6
6
4
1
12
m

b2
þ c2


1
12
mab 
1
12
mac

1
12
mab
1
12
m

a2
þ c2


1
12
mbc

1
12
mac 
1
12
mbc
1
12
m

a2
þ b2

3
7
7
7
7
7
7
7
7
7
7
5
(9.63)
FIGURE 9.15
A uniform slender rod.
492 CHAPTER 9 Rigid Body Dynamics
EXAMPLE 9.11
Calculate the principal moments of inertia about the center of mass and the corresponding principal directions for
the bent rod in Figure 9.16. Its mass is uniformly distributed at 2 kg/m.
Solution
The mass of each of the rod segments is
m1 ¼ 2$0:4 ¼ 0:8 kg m2 ¼ 2$0:5 ¼ 1 kg m3 ¼ 2$0:3 ¼ 0:6 kg m4 ¼ 2$0:2 ¼ 0:4 kg (a)
The total mass of the system is
m ¼
X
4
i¼1
mi ¼ 2:8 kg (b)
The coordinates of each segment’s center of mass are
xG1
¼ 0 yG1
¼ 0 zG1
¼ 0:2 m
xG2
¼ 0 yG2
¼ 0:25 m zG2
¼ 0:2 m
xG3
¼ 0:15 m yG3
¼ 0:5 m zG3
¼ 0
xG4
¼ 0:3 m yG4
¼ 0:4 m zG4
¼ 0
(c)
If the slender rod in Figure 9.15 is aligned with, say, the x-axis, then a ¼ [ and b ¼ c ¼ 0, so that according to
Eqn (9.63),
½IG ¼
2
6
6
6
6
6
4
0 0 0
0
1
12
m‘2
0
0 0
1
12
m‘2
3
7
7
7
7
7
5
That is, the moment of inertia of a slender rod about the axes normal to the rod at its center of mass is 1
12 m‘2,
where m and [ are the mass and length of the rod, respectively. Since the mass of a slender bar is assumed to be
concentrated along the axis of the bar (its cross-sectional dimensions are infinitesimal), the moment of inertia
4
3
2
1
0.4 m
0.5 m
0.3 m
0.2 m
O
FIGURE 9.16
Bent rod for which the principal moments of inertia are to be determined.
9.5 Moments of inertia 493
about the centerline is zero. By symmetry, the products of inertia about the axes through the center of mass are all
zero. Using this information and the parallel axis theorem, we find the moments and products of inertia of each rod
segment about the origin O of the xyz system as follows:
Rod 1:
I
ð1Þ
x ¼ I
ð1Þ
G1

x
þ m1

y2
G1
þ z2
G1

¼

1
12
$0:8$0:42

þ
h
0:8

0 þ 0:22
i
¼ 0:04267 kg-m2
I
ð1Þ
y ¼ I
ð1Þ
G1

y
þ m1

x2
G1
þ z2
G1

¼

1
12
$0:8$0:42

þ
h
0:8

0 þ 0:22
i
¼ 0:04267 kg-m2
I
ð1Þ
z ¼ I
ð1Þ
G1

z
þ m1

x2
G1
þ y2
G1

¼ 0 þ ½0:8ð0 þ 0Þ ¼ 0
I
ð1Þ
xy ¼ I
ð1Þ
G1

xy
 m1xG1
yG1
¼ 0  ½0:8ð0Þð0Þ ¼ 0
I
ð1Þ
xz ¼ I
ð1Þ
G1

xz
 m1xG1
zG1
¼ 0  ½0:8ð0Þð0:2Þ ¼ 0
I
ð1Þ
yz ¼ I
ð1Þ
G1

yz
 m1yG1
zG1
¼ 0  ½0:8ð0Þð0Þ ¼ 0
Rod 2:
I
ð2Þ
x ¼ I
ð2Þ
G2

x
þ m2

y2
G2
þ z2
G2

¼

1
12
$1:0$0:52

þ
h
1:0

0 þ 0:252
i
¼ 0:08333 kg-m2
I
ð2Þ
y ¼ I
ð2Þ
G2

y
þ m2

x2
G2
þ z2
G2

¼ 0 þ ½1:0ð0 þ 0Þ ¼ 0
I
ð2Þ
z ¼ I
ð2Þ
G2

z
þ m2

x2
G2
þ y2
G2

¼

1
12
$1:0$0:52

þ
h
1:0

0 þ 0:52
i
¼ 0:08333 kg-m2
I
ð2Þ
xy ¼ I
ð2Þ
G2

xy
 m2xG2
yG2
¼ 0  ½1:0ð0Þð0:5Þ ¼ 0
I
ð2Þ
xz ¼ I
ð2Þ
G2

xz
 m2xG2
zG2
¼ 0  ½1:0ð0Þð0Þ ¼ 0
I
ð2Þ
yz ¼ I
ð2Þ
G2

yz
 m2yG2
zG2
¼ 0  ½1:0ð0:5Þð0Þ ¼ 0
Rod 3:
I
ð3Þ
x ¼ I
ð3Þ
G3

x
þ m3

y2
G3
þ z2
G3

¼ 0 þ 0:6

0:52 þ 0

¼ 0:15 kg-m2
I
ð3Þ
y ¼ I
ð3Þ
G3

y
þ m3

x2
G3
þ z2
G3

¼

1
12
$0:6$0:32

þ
h
0:6

0:152
þ 0
i
¼ 0:018 kg-m2
I
ð3Þ
z ¼ I
ð3Þ
G3

z
þ m3

x2
G3
þ y2
G3

¼

1
12
$0:6$0:32

þ
h
0:6

0:152
þ 0:52
i
¼ 0:1680 kg-m2
I
ð3Þ
xy ¼ I
ð3Þ
G3

xy
 m3xG3
yG3
¼ 0  ½0:6ð0:15Þð0:5Þ ¼ 0:045 kg-m2
I
ð3Þ
xz ¼ I
ð3Þ
G3

xz
 m3xG3
zG3
¼ 0  ½0:6ð0:15Þð0Þ ¼ 0
I
ð3Þ
yz ¼ I
ð3Þ
G3

yz
 m3yG3
zG3
¼ 0  ½0:6ð0:5Þð0Þ ¼ 0
494 CHAPTER 9 Rigid Body Dynamics
Rod 4:
I
ð4Þ
x ¼ I
ð4Þ
G4

x
þ m4

y2
G4
þ z2
G4

¼

1
12
$0:4$0:22

þ
h
0:4

0:42
þ 0
i
¼ 0:06533 kg-m2
I
ð4Þ
y ¼ I
ð4Þ
G4

y
þ m4

x2
G4
þ z2
G4

¼ 0 þ 0:4

0:32 þ 0

¼ 0:0360 kg-m2
I
ð4Þ
z ¼ I
ð4Þ
G4

z
þ m4

x2
G4
þ y2
G4

¼

1
12
$0:4$0:22

þ
h
0:4

0:32
þ 0:42
i
¼ 0:1013 kg-m2
I
ð4Þ
xy ¼ I
ð4Þ
G4

xy
 m4xG4
yG4
¼ 0  ½0:4ð0:3Þð0:4Þ ¼ 0:0480 kg-m2
I
ð4Þ
xz ¼ I
ð4Þ
G4

xz
 m4xG4
zG4
¼ 0  ½0:4ð0:3Þð0Þ ¼ 0
I
ð4Þ
yz ¼ I
ð4Þ
G4

yz
 m4yG4
zG4
¼ 0  ½0:4ð0:4Þð0Þ ¼ 0
The total moments of inertia for all the four rods about O are
Ix ¼
P
4
i¼1
I
ðiÞ
x ¼ 0:3413 kg-m2
Iy ¼
P
4
i¼1
I
ðiÞ
y ¼ 0:09667 kg-m2
Iz ¼
P
4
i¼1
I
ðiÞ
z ¼ 0:3527 kg-m2
Ixy ¼
P
4
i¼1
I
ðiÞ
xy ¼ 0:0930 kg-m2
Ixz ¼
P
4
i¼1
I
ðiÞ
xz ¼ 0 Iyz ¼
P
4
i¼1
I
ðiÞ
yz ¼ 0
(d)
The coordinates of the center of mass of the system of four rods are, from Eqns (a)e(c),
xG ¼
1
m
X
4
i¼1
mixGi
¼
1
2:8
$0:21 ¼ 0:075 m
yG ¼
1
m
X
4
i¼1
miyGi
¼
1
2:8
$0:71 ¼ 0:2536 m
zG ¼
1
m
X
4
i¼1
mizGi
¼
1
2:8
$0:16 ¼ 0:05714 m
(e)
We use the parallel axis theorems to shift the moments of inertia in Eqn (d) to the center of mass G of the system.
IGx
¼ Ix  m

y2
G þ z2
G

¼ 0:3413  0:1892 ¼ 0:1522 kg-m2
IGy
¼ Iy  m

x2
G þ z2
G

¼ 0:09667  0:02489 ¼ 0:07177 kg-m2
IGz
¼ Iz  m

x2
G þ y2
G

¼ 0:3527  0:1958 ¼ 0:1569 kg-m2
IGxy
¼ Ixy þ mxGyG ¼ 0:093 þ 0:05325 ¼ 0:03975 kg-m2
IGxz
¼ Ixz þ mxGzG ¼ 0 þ 0:012 ¼ 0:012 kg-m2
IGyz
¼ Iyz þ myGzG ¼ 0 þ 0:04057 ¼ 0:04057 kg-m2
9.5 Moments of inertia 495
Therefore, the inertia tensor, relative to the center of mass, is
½I ¼
2
6
6
4
IGx
IGxy
IGxz
IGxy
IGy
IGyz
IGxz
IGyz
IGz
3
7
7
5 ¼
2
6
6
4
0:1522 0:03975 0:012
0:03975 0:07177 0:04057
0:012 0:04057 0:1569
3
7
7
5

kg$m2

(f)
To find the three principal moments of inertia, we may proceed as in Example 9.9, or simply enter the following
lines in the MATLAB Command Window:
I ¼ [ 0.1522 -0.03975 0.012
-0.03975 0.07177 0.04057
0.012 0.04057 0.1569 ];
[eigenVectors, eigenValues] ¼ eig(IG)
to obtain
eigenVectors ¼
0.3469 -0.8482 -0.4003
0.8742 0.1378 0.4656
-0.3397 -0.5115 0.7893
eigenValues ¼
0.0402 0 0
0 0.1658 0
0 0 0.1747
Hence, the three principal moments of inertia and their principal directions are
l1 ¼ 0:04023 kg-m2
eð1Þ ¼ 0:3469^
i þ 0:8742^
j  0:3397^
k
l2 ¼ 0:1658 kg-m2
eð2Þ ¼ 0:8482^
i þ 0:1378^
j  0:5115^
k
l3 ¼ 0:1747 kg-m2
eð3Þ ¼ 0:4003^
i þ 0:4656^
j þ 0:7893^
k
9.6 Euler’s equations
For either the center of mass G or for a fixed point P about which the body is in pure rotation, we know
from Eqns (9.29) and (9.30) that
Mnet ¼ _
H (9.64)
Using a comoving coordinate system, with angular velocity U and its origin located at the point (G or
P), the angular momentum has the analytical expression
H ¼ Hx
^
i þ Hy
^
j þ Hz
^
k (9.65)
We shall henceforth assume, for simplicity, that
ðaÞ The moving xyz axes are the principal axes of inertia and (9.66a)
ðbÞ The moments of inertia relative to xyz are constant in time: (9.66b)
496 CHAPTER 9 Rigid Body Dynamics
Equations (9.42) and (9.66a) imply that
H ¼ Aux
^
i þ Buy
^
j þ Cuz
^
k (9.67)
where A, B, and C are the principal moments of inertia.
According to Eqn (1.56), the time derivative of H is _
H ¼ _
HÞrel þ U  H, so that Eqn (9.64) can be
written as
Mnet ¼ _
H

rel
þ U  H (9.68)
Keep in mind that, whereas U (the angular velocity of the moving xyz coordinate system) and u (the
angular velocity of the rigid body itself) are both absolute kinematic quantities, Eqn (9.68) contains
their components as projected onto the axes of the noninertial xyz frame given by
u ¼ ux
^
i þ uy
^
j þ uz
^
k
U ¼ Ux
^
i þ Uy
^
j þ Uz
^
k
The absolute angular acceleration a is obtained using Eqn (1.56) as
a ¼ _
u ¼
dux
dt
^
i þ
duy
dt
^
j þ
duz
dt
^
k
zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{
arel
þ U  u
that is,
a ¼

_
ux þ Uyuz  Uzuy

^
i þ

_
uy þ Uzux  Uxuz

^
j þ

_
uz þ Uxuy  Uyux

^
k (9.69)
Clearly, it is generally true that
axs _
ux ays _
uy azs _
uz
From Eqn (1.57) and Eqn (9.67),
_
H

rel
¼
dðAuxÞ
dt
^
i þ
d

Buy

dt
^
j þ
dðCuzÞ
dt
^
k
Since A, B, and C are constant, this becomes
_
H

rel
¼ A _
ux
^
i þ B _
uy
^
j þ C _
uz
^
k (9.70)
Substituting Eqns (9.67) and (9.70) into Eqn (9.68) yields
Mnet ¼ A _
ux
^
i þ B _
uy
^
j þ C _
uz
^
k þ
^
i ^
j ^
k
Ux Uy Uz
Aux Buy Cuz
Expanding the crossproduct and collecting the terms leads to
Mxnet
¼ A _
ux þ CUyuz  BUzuy
Mynet
¼ B _
uy þ AUzux  CUxuz
Mznet
¼ C _
uz þ BUxuy  AUyux
(9.71)
9.6 Euler’s equations 497
If the comoving frame is a rigidly attached body frame, then its angular velocity is the same as that
of the body, that is, U ¼ u. In that case, Eqn (9.68) reduces to the classical Euler’s equation of motion,
namely,
Mnet ¼ _
H

rel
þ u  H (9.72a)
the three components of which are obtained from Eqn (9.71) as
Mxnet
¼ A _
ux þ ðC  BÞuyuz
Mynet
¼ B _
uy þ ðA  CÞuzux
Mznet
¼ C _
uz þ ðB  AÞuxuy
(9.72b)
Equation (9.68) is sometimes referred to as the modified Euler equation.
When U ¼ u, it follows from Eqn (9.69) that
_
ux ¼ ax _
uy ¼ ay _
uz ¼ az (9.73)
That is, the relative angular acceleration equals the absolute angular acceleration when U ¼ u. Rather
than calculating the time derivatives _
ux, _
uy, and _
uz for use in Eqn (9.72), we may in this case first
compute a in the absolute XYZ frame
a ¼
du
dt
¼
duX
dt
^
I þ
duY
dt
^
J þ
duZ
dt
^
K
and then project these components onto the xyz body frame, so that
8

:
_
ux
_
uy
_
uz
9
=
;
¼ ½QXx
8

:
duX=dt
duY=dt
duZ=dt
9
=
;
(9.74)
where [Q]Xx is the time-dependent orthogonal transformation from the inertial XYZ frame to the
noninertial xyz frame.
EXAMPLE 9.12
Calculate the net moment on the solar panel of Examples 9.2 and 9.8 (Figure 9.17).
Solution
Since the comoving frame is rigidly attached to the panel, Euler’s equation (Eqn (9.72a)) applies to this problem,
that is
MGnet
¼ _
HG

rel
þ u  HG (a)
where
HG ¼ Aux
^
i þ Buy
^
j þ Cuz
^
k (b)
and
_
HG

rel
¼ A _
ux
^
i þ B _
uy
^
j þ C _
uz
^
k (c)
498 CHAPTER 9 Rigid Body Dynamics
In Example 9.2, the angular velocity of the panel in the satellite’s x0y0z0 frame was found to be
u ¼ _
q^
j
0
þ N^
k
0
(d)
In Example 9.8, we showed that the transformation from the panel’s xyz frame to that of the satellite is represented
by the matrix
½Q ¼
2
6
6
4
sin q 0 cos q
0 1 0
cos q 0 sin q
3
7
7
5 (e)
We use the transpose of [Q] to transform the components of u into the panel frame of reference,
fugxyz ¼ ½QT
fugx0y0z0 ¼
2
6
6
4
sin q 0 cos q
0 1 0
cos q 0 sin q
3
7
7
5
8





:
0
_
q
N
9


=


;
¼
8





:
N cos q
_
q
N sin q
9


=


;
or
ux ¼ N cos q uy ¼ _
q uz ¼ N sin q (f)
In Example 9.2, N and _
q were said to be constant. Therefore, the time derivatives of Eqn (f) are
_
ux ¼
dðN cos qÞ
dt
¼ N _
q sin q
_
uy ¼
d_
q
dt
¼ 0
_
uz ¼
dðN sin qÞ
dt
¼ N _
q cos q
(g)
In Example 9.8, the moments of inertia in the panel frame of reference were listed as
A ¼
1
12
m

‘2
þ t2

B ¼
1
12
m

w2
þ t2

C ¼
1
12
m

w2
þ ‘2


IGxy ¼ IGxz ¼ IGyz ¼ 0

(h)
θ
FIGURE 9.17
Free-body diagram of the solar panel in Examples 9.2 and 9.8.
9.6 Euler’s equations 499
Substituting Eqns (b), (c), (f), (g) and (h) into Eqn (a) yields
MGnet
¼
1
12
m

‘2
þ t2

N _
q sin q
^
i þ
1
12
m

w2
þ t2

$0$^
j þ
1
12
m

w2
þ ‘2

N _
q cos q

^
k
þ
^
i ^
j ^
k
N cos q _
q N sin q
1
12
m

‘2
þ t2

ðN cos qÞ
1
12
m

w2
þ t2

_
q
1
12
m

w2
þ ‘2

ðN sin qÞ
Upon expanding the cross product and collecting terms, this reduces to
MGnet
¼ 
1
6
mt2
N _
q sin q^
i þ
1
24
m

t2
 w2

N2
sin 2q^
j þ
1
6
mw2
N _
q cos q^
k
Using the numerical data of Example 9.8 (m ¼ 50 kg, N ¼ 0.1 rad/s, q ¼ 40, _
q ¼ 0:01 rad=s, [ ¼ 6 m, w ¼ 2 m,
and t ¼ 0.025 m), we find
MGnet
¼ 3:348  106^
i  0:08205^
j þ 0:02554^
k ðN$mÞ
EXAMPLE 9.13
Calculate the net moment on the gyro rotor of Examples 9.3 and 9.6.
Solution
Figure 9.18 is a free-body diagram of the rotor. Since in this case the comoving frame is not rigidly attached to the
rotor, we must use Eqn (9.68) to find the net moment about G, that is
MGnet
¼ _
HG

rel
þ U  HG (a)
where
HG ¼ Aux
^
i þ Buy
^
j þ Cuz
^
k (b)
and
_
HG

rel
¼ A _
ux
^
i þ B _
uy
^
j þ C _
uz
^
k (c)
From Eqn (i) of Example 9.3, we know that the components of the angular velocity of the rotor in the moving
reference frame are
ux ¼ _
q
uy ¼ N sin q
uz ¼ uspin þ N cos q
(d)
500 CHAPTER 9 Rigid Body Dynamics
Since, as specified in Example 9.3, _
q, N, and uspin are all constant, it follows that
_
ux ¼
d_
q
dt
¼ 0
_
uy ¼
dðN sin qÞ
dt
¼ N _
q cos q
_
uz ¼
d

uspin þ N cos q

dt
¼ N _
q sin q
(e)
The angular velocity U of the comoving xyz frame is that of the gimbal ring, which equals the angular velocity of
the rotor minus its spin. Therefore,
Ux ¼ _
q
Uy ¼ N sin q
Uz ¼ N cos q
(f)
In Example 9.6, we found that
A ¼ B ¼
1
12
mt2
þ
1
4
mr2
C ¼
1
2
mr2
(g)
FIGURE 9.18
Free-body diagram of the gyro rotor of Examples 9.6 and 9.3.
9.6 Euler’s equations 501
Substituting Eqns (b) through (g) into Eqn (a), we get
MGnet
¼

1
12
mt2
þ
1
4
mr2

$0^
i þ

1
12
mt2
þ
1
4
mr2


N _
q cos q
^
j þ
1
2
mr2

N _
q sin q

^
k
þ
^
i ^
j ^
k
_
q N sin q N cos q

1
12
mt2
þ
1
4
mr2

_
q

1
12
mt2
þ
1
4
mr2

N sin q
1
2
mr2

uspin þ N cos q

Expanding the cross product, collecting terms, and simplifying leads to
MGnet
¼

1
2
uspin þ
1
12

3 
t2
r2

N cos q mr2
N sin q^
i þ

1
6
t2
r2
N cos q 
1
2
uspin

mr2 _
q^
j 
1
2
N _
q sin qmr2^
k (h)
In Example 9.3, the following numerical data were provided: m ¼ 5 kg, r ¼ 0.08 m, t ¼ 0.025 m, N ¼ 2.1 rad/s,
q ¼ 60, _
q ¼ 4 rad=s, and uspin ¼ 105 rad/s. For this set of numbers, Eqn (h) becomes
MGnet
¼ 0:3203^
i  0:6698^
j  0:1164^
k ðN$mÞ
9.7 Kinetic energy
The kinetic energy T of a rigid body is the integral of the kinetic energy 1
2 v2dm of its individual mass
elements,
T ¼
Z
m
1
2
v2
dm ¼
Z
m
1
2
v$vdm (9.75)
where v is the absolute velocity _
R of the element of mass dm. From Figure 9.8, we infer that
_
R ¼ _
RG þ _
r. Furthermore, Eqn (1.52) requires that _
r ¼ u  r. Thus, v ¼ vG þ u  r, which means
v$v ¼ ½vG þ u  r$½vG þ u  r ¼ vG
2
þ 2vG$ðu  rÞ þ ðu  rÞ$ðu  rÞ
We can apply the vector identity introduced in Eqn (1.21), namely
A$ðB  CÞ ¼ B$ðC  AÞ (9.76)
to the last term to get
v$v ¼ vG
2
þ 2vG$ðu  rÞ þ u$½r  ðu  rÞ
Therefore, Eqn (9.75) becomes
T ¼
Z
m
1
2
vG
2
dm þ vG$
0
@u 
Z
m
rdm
1
A þ
1
2
u$
Z
m
r  ðu  rÞdm
502 CHAPTER 9 Rigid Body Dynamics
Since r is measured from the center of mass,
R
m
rdm ¼ 0. Recall that, according to Eqn (9.34),
Z
m
r  ðu  rÞdm ¼ HG
It follows that the kinetic energy may be written as
T ¼
1
2
mvG
2
þ
1
2
u$HG (9.77)
The second term is the rotational kinetic energy TR,
TR ¼
1
2
u$HG (9.78)
If the body is rotating about a point P that is at rest in inertial space, we have from Eqn (9.2) and
Figure 9.8 that
vG ¼ vP þ u  rG=P ¼ 0 þ u  rG=P ¼ u  rG=P
It follows that
vG
2
¼ vG$vG ¼

u  rG=P

$

u  rG=P

Making use once again of the vector identity in Eqn (9.76), we find
vG
2
¼ u$
h
rG=P 

u  rG=P
i
¼ u$

rG=P  vG

Substituting this into Eqn (9.77) yields
T ¼
1
2
u$
h
HG þ rG=P  mvG
i
Equation (9.21) shows that this can be written as
T ¼
1
2
u$HP (9.79)
In this case, of course, all of the kinetic energy is rotational.
In terms of the components of u and H, whether it is HP or HG, the rotational kinetic energy
expression becomes, with the aid of Eqn (9.39),
TR ¼
1
2

uxHx þ uyHy þ uyHz

¼
1
2
P ux uy uz R
2
6
6
4
Ix Ixy Ixz
Ixy Iy Iyz
Ixz Iyz Iz
3
7
7
5
8





:
ux
uy
uz
9


=


;
Expanding, we obtain
TR ¼
1
2
Ixux
2
þ
1
2
Iyuy
2
þ
1
2
Izuz
2
þ Ixyuxuy þ Ixzuxuz þ Iyzuyuz (9.80)
9.7 Kinetic energy 503
Obviously, if the xyz axes are principal axes of inertia, then Eqn (9.80) simplifies considerably,
TR ¼
1
2
Aux
2
þ
1
2
Buy
2
þ
1
2
Cuz
2
(9.81)
EXAMPLE 9.14
A satellite in a circular geocentric orbit of 300-km altitude has a mass of 1500 kg, and the moments of inertia
relative to a body frame with origin at the center of mass G are
½I ¼
2
6
6
4
2000 1000 2500
1500 3000 1500
2500 1500 4000
3
7
7
5

kg$m2

If at a given instant the components of angular velocity in this frame of reference are
u ¼ 1^
i  0:9^
j þ 1:5^
k ðrad=sÞ
calculate the total kinetic energy of the satellite.
Solution
The speed of the satellite in its circular orbit is
v ¼
ffiffiffi
m
r
r
¼
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
398; 600
6378 þ 300
r
¼ 7:7258 km=s
The angular momentum of the satellite is
fHGg ¼ ½IGfug ¼
2
6
6
4
2000 1000 2500
1500 3000 1500
2500 1500 4000
3
7
7
5
8





:
1
0:9
1:5
9


=


;
¼
8





:
6650
5950
9850
9


=


;

kg$m2
=s

Therefore, the total kinetic energy is
T ¼
1
2
mvG
2
þ
1
2
u$HG ¼
1
2
$1500$7725:82
þ
1
2
P 1 0:9 1:5 R
8





:
6650
5950
9850
9


=


;
¼ 44:766  106
þ 13; 390
T ¼ 44:766 MJ
Obviously, the kinetic energy is dominated by that due to the orbital motion.
9.8 The spinning top
Let us analyze the motion of the simple axisymmetric top in Figure 9.19. It is constrained to rotate
about point O.
The moving coordinate system is chosen to have its origin at O. The z-axis is aligned with the spin
axis of the top (the axis of rotational symmetry). The x-axis is the node line, which passes through O
504 CHAPTER 9 Rigid Body Dynamics
and is perpendicular to the plane defined by the inertial Z-axis and the spin axis of the top. The y-axis
is then perpendicular to x and z, such that ^
j ¼ ^
k ^
i. By symmetry, the moment of inertia matrix of the
top relative to the xyz frame is diagonal, with Ix ¼ Iy ¼ A and Iz ¼ C. From Eqns (9.68) and (9.70),
we have
M0net
¼ A _
ux
^
i þ A _
uy
^
j þ C _
uz
^
k þ
^
i ^
j ^
k
Ux Uy Uz
Aux Auy Cuz
(9.82)
The angular velocity u of the top is the vector sum of the spin rate us and the rates of precession up and
nutation un, where
up ¼ _
f un ¼ _
q (9.83)
Thus,
u ¼ un
^
i þ up
^
K þ us
^
k
From the geometry, we see that
^
K ¼ sin q^
j þ cos q^
k (9.84)
Therefore, relative to the comoving system,
u ¼ un
^
i þ up sin q^
j þ

us þ up cos q

^
k (9.85)
FIGURE 9.19
Simple top rotating about the fixed point O.
9.8 The spinning top 505
From Eqn (9.85), we see that
ux ¼ un uy ¼ up sin q uz ¼ us þ up cos q (9.86)
Computing the time rates of these three expressions yields the components of angular acceleration
relative to the xyz frame, given by
_
ux ¼ _
un _
uy ¼ _
up sin q þ upun cos q _
uz ¼ _
us þ _
up cos q  upun sin q (9.87)
The angular velocity U of the xyz system is U ¼ up
^
K þ un
^
i, so that, using Eqn (9.84),
U ¼ un
^
i þ up sin q^
j þ up cos q^
k (9.88)
From Eqn (9.88), we obtain
Ux ¼ un Uy ¼ up sin q Uz ¼ up cos q (9.89)
In Figure 9.19, the moment about O is that of the weight vector acting through the center of mass G:
M0net
¼

d^
k



mg^
K

¼ mgd^
k 

sin q^
j þ cos q^
k

or
M0net
¼ mgd sin q^
i (9.90)
Substituting Eqns (9.86), (9.87), (9.89), and (9.90) into Eqn (9.82), we get
mgd sin q^
i ¼ A _
un
^
i þ A

_
up sin q þ upun cos q

^
j þ C

_
us þ _
up cos q  upun sin q

^
k
þ
^
i ^
j ^
k
un up sin q up cos q
Aun Aup sin q C

us þ up cos q

(9.91)
Let us consider the special case in which q is constant, that is, there is no nutation, so that
un ¼ _
un ¼ 0. Then, Eqn (9.91) reduces to
mgd sin q^
i ¼ A _
up sin q^
j þ C

_
us þ _
up cos q

^
k þ
^
i ^
j ^
k
0 up sin q up cos q
0 Aup sin q C

us þ up cos q

(9.92)
Expanding the determinant yields
mgd sin q^
i ¼ A _
up sin q^
j þ C

_
us þ _
up cos q

^
k þ
h
Cupus sin q þ ðC  AÞup
2
cos q sin q
i
^
i
Equating the coefficients of^
i,^
j, and ^
k on each side of the equation and assuming that 0  q  180
leads to
mgd ¼ Cupus þ ðC  AÞup
2
cos q (9.93a)
506 CHAPTER 9 Rigid Body Dynamics
A _
up ¼ 0 (9.93b)
C

_
us þ _
up cos q

¼ 0 (9.93c)
Equation (9.93b) implies _
up ¼ 0, and from Eqn (9.93c), it follows that _
us ¼ 0. Therefore, the rates
of spin and precession are both constant. From Eqn (9.93a), we find
ðA  CÞcos qup
2
 Cusup þ mgd ¼ 0 (9.94)
If the spin rate is zero, Eqn (9.94) yields
up

us¼0
¼ 
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
mgd
ðC  AÞcos q
s
if ðC  AÞ cos q  0 (9.95)
In this case, the top rotates about O at this rate, without spinning. If A  C (prolate), its symmetry axis
must make an angle between 90 and 180 to the vertical; otherwise up is imaginary. On the other
hand, if A  C (oblate), the angle lies between 0 and 90. Thus, in steady rotation without spin, the
top’s axis sweeps out a cone that lies either below the horizontal plane (A  C) or above the plane
(A  C).
In the special case where (A  C) cos q ¼ 0, Eqn (9.94) yields a steady precession rate that is
inversely proportional to the spin rate,
up ¼
mgd
Cus
if ðA  CÞ cos q ¼ 0 (9.96)
If A ¼ C, this precession apparently occurs irrespective of tilt angle q. If A s C, this rate of precession
occurs at q ¼ 90, that is, the spin axis is perpendicular to the precession axis.
In general, Eqn (9.94) is a quadratic equation in up, so we can use the quadratic formula
to find
up ¼
C
2ðA  CÞcos q
us 
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
us
2 
4mgdðA  CÞcos q
C2
r !
(9.97)
Thus, for a given spin rate and tilt angle q (q s 90), there are two rates of precession _
f.
Observe that if (A  C)cos q  0, then up is imaginary when us
2  4mgdðA  CÞcos q=C2.
Therefore, the minimum spin rate required for steady precession at a constant inclination q is
usÞmin ¼
2
C
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
mgdðA  CÞcos q
p
if ðA  CÞcos q  0 (9.98)
If (A  C) cos q  0, the radical in Eqn (9.97) is real for all us. In this case, as us / 0, up approaches
the value given above in Eqn (9.95).
9.8 The spinning top 507
EXAMPLE 9.15
Calculate the precession rate up for the top of Figure 9.19 if m ¼ 0.5 kg, A (¼ Ix ¼ Iy) ¼ 12  104
kg m2
, C
(¼ Iz) ¼ 4.5  104
kg m2
, and d ¼ 0.05 m.
Solution
For an inclination of, say, 60, (A  C) cos q  0, so that Eqn (9.98) requires usÞmin ¼ 407:01 rpm. Let us choose
the spin rate to be us ¼ 1000 rpm ¼ 104.7 rad/sec. Then, from Eqn (9.97), the precession rate as a function of the
inclination q is given by either one of the following formulas:
up ¼ 31:42
1 þ
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1  0:3312 cos q
p
cos q
and up ¼ 31:42
1 
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1  0:3312 cos q
p
cos q
(a)
These are plotted in Figure 9.20.
Figure 9.21 shows an axisymmetric rotor mounted so that its spin axis (z) remains perpendicular to the
precession axis (y). In that case, Eqn (9.85) with q ¼ 90 yields
u ¼ up
^
j þ us
^
k (9.99)
Likewise, from Eqn (9.88), the angular velocity of the comoving xyz system is U ¼ up
^
j. If we assume
that the spin rate and precession rate are constant ðdup=dt ¼ dus=dt ¼ 0Þ, then Eqn (9.68), written for
the center of mass G, becomes
MGnet
¼ U  H ¼

up
^
j



Aup
^
j þ Cus
^
k

(9.100)
where A and C are the moments of inertia of the rotor about the x- and z-axes, respectively. Setting
Cus
^
k ¼ Hs, the spin angular momentum, and up
^
j ¼ up, we obtain
MGnet
¼ up  Hs

Hs ¼ Cus
^
k

(9.101)
Since the center of mass is the reference point, there is no restriction on the motion G for which Eqn
(9.101) is valid. Observe that the net gyroscopic moment MGnet
exerted on the rotor by its supports is
90
ω
180
–6000
–3000
0
6000
45
50
55
θ (°)
180
90
0
0
3000
p (rpm) ωp (rpm)
θ (°)
45 135 45 135
(a) (b)
FIGURE 9.20
(a) High-energy precession rate (unlikely to be observed). (b) Low energy precession rate (the one most always
seen).
508 CHAPTER 9 Rigid Body Dynamics
perpendicular to the plane of the spin and the precession vectors. If a spinning rotor is forced to
precess, the gyroscopic moment MGnet
develops. Or, if a moment is applied normal to the spin axis of a
rotor, it will precess so as to cause the spin axis to turn toward the moment axis.
EXAMPLE 9.16
A uniform cylinder of radius r, length L, and mass m spins at a constant angular velocity us. It rests on simple
supports (which cannot exert couples), mounted on a platform that rotates at an angular velocity of up. Find the
reactions at A and B. Neglect the weight (i.e., calculate the reactions due just to the gyroscopic effects).
Solution
The net vertical force on the cylinder is zero, so the reactions at each end must be equal and opposite in direction,
as shown on the free-body diagram insert in Figure 9.22. Noting that the moment of inertia of a uniform cylinder
about it axis of rotational symmetry is 1
2mr2, Eqn (9.101) yields
RL^
i ¼

up
^
j



1
2
mr2
us
^
k

¼
1
2
mr2
up us
^
i
so that
R ¼
mr2upus
2L
ω
ω
FIGURE 9.22
Illustration of the gyroscopic effect.
ω
ω
FIGURE 9.21
A spinning rotor on a rotating platform.
9.8 The spinning top 509
9.9 Euler angles
Three angles are required to specify the orientation of a rigid body relative to an inertial frame.
The choice is not unique, but there are two sets in common use: the Euler angles and the yaw,
pitch, and roll angles. We will discuss each of them in turn. The reader is urged to review Section
4.5 on orthogonal coordinate transformations and, in particular, the discussion of Euler angle
sequences.
The three Euler angles f, q, and j shown in Figure 9.23 give the orientation of a body-
fixed xyz frame of reference relative to the XYZ inertial frame of reference. The xyz frame is
obtained from the XYZ frame by a sequence of rotations through each of the Euler angles in
turn. The first rotation is around the Z (¼ z1) axis through the precession angle f. This takes
X into x1 and Y into y1. The second rotation is around the x2 (¼ x1) axis through the
nutation angle q. This carries y1 and z1 into y2 and z2, respectively. The third and final
rotation is around the z (¼ z2) axis through the spin angle j, which takes x2 into x and y2
into y.
The matrix [Q]Xx of the transformation from the inertial frame to the body-fixed frame is given by
the classical Euler angle sequence (Eqn (4.37)):
½QXx ¼ ½R3ðjÞ½R1ðqÞ½R3ðfÞ (9.102)
2
1
3
3
2
1
X
Y
x1 x2
,
Z z1
,
y1
x
y
y2
z2, z
φ
φ
φ
θ ψ
ψ
ψ
θ
θ
FIGURE 9.23
Classical Euler angle sequence. See also Figure 4.14.
510 CHAPTER 9 Rigid Body Dynamics
From Eqns (4.32) and (4.34), we have
½R3ðjÞ ¼
2
6
6
6
4
cos j sin j 0
sin j cos j 0
0 0 1
3
7
7
7
5
½R1ðqÞ ¼
2
6
6
6
4
1 0 0
0 cos q sin q
0 sin q cos q
3
7
7
7
5
½R3ðfÞ ¼
2
6
6
6
4
cos f sin f 0
sin f cos f 0
0 0 1
3
7
7
7
5
(9.103)
According to Eqn (4.38), the direction cosine matrix is
½QXx ¼
2
6
6
4
sin f cos q sin j þ cos f cos j cos f cos q sin j þ sin f cos j sin q sin j
sin f cos q cos g  cos f sin j cos f cos q cos j  sin f sin j sin q cos j
sin f sin q cos f sin q cos q
3
7
7
5
(9.104)
Since this is an orthogonal matrix, the inverse transformation from xyz to XYZ is ½QxX ¼ ð½QXxÞT
,
½QxX ¼
2
6
6
4
sin f cos q sin j þ cos f cos j sin f cos q cos g  cos f sin j sin f sin q
cos f cos q sin j þ sin f cos j cos f cos q cos j  sin f sin j cos f sin q
sin q sin j sin q cos j cos q
3
7
7
5
(9.105)
Algorithm 4.3 is used to find the three Euler angles q, f, and j from a given direction cosine
matrix [Q]Xx.
EXAMPLE 9.17
The direction cosine matrix of an orthogonal transformation from XYZ to xyz is
½Q ¼
2
6
6
4
0:32175 0:89930 0:29620
0:57791 0:061275 0:81380
0:75000 0:43301 0:5000
3
7
7
5
Use Algorithm 4.3 to find the Euler angles f, q, and j for this transformation.
Solution
Step1 (precession angle):
f ¼ tan1

Q31
Q32

¼ tan1

0:75000
0:43301

ð0  f  360
Þ
9.9 Euler angles 511
Since the numerator is negative and the denominator is positive, the angle f lies in the fourth quadrant.
f ¼ tan1
ð1:7320Þ ¼ 300
Step 2 (nutation angle):
q ¼ cos1
Q33 ¼ cos1
ð0:5000Þ ð0  q  180
Þ
q ¼ 120
Step 3 (spin angle):
j ¼ tan1Q13
Q23
¼ tan1

0:29620
0:81380

ð0  j  360
Þ
Since both the numerator and denominator are negative, the angle j lies in the third quadrant.
j ¼ tan1
ð0:36397Þ ¼ 200
The time rates of change of the Euler angles f, q, and j are, respectively, the precession rate up, the
nutation rate un, and the spin us. That is,
up ¼ _
f un ¼ _
q us ¼ _
j (9.106)
The absolute angular velocity u of a rigid body can be resolved into components ux, uy, and uz along
the body-fixed xyz axes, so that
u ¼ ux
^
i þ uy
^
j þ uz
^
k (9.107)
Figure 9.23 shows that precession is measured around the inertial Z-axis (unit vector ^
K); nutation is
measured around the intermediate x1-axis (node line) with unit vector^
i1; and spin is measured around
the body-fixed z-axis (unit vector ^
k). Therefore, the absolute angular velocity can alternatively be
written in terms of the nonorthogonal Euler angle rates as
u ¼ up
^
K þ un
^
i1 þ us
^
k (9.108)
In order to find the relationship between the body rates ux, uy, and uz and the Euler angle rates up, un,
and us, we must express ^
K and ^
i1 in terms of the unit vectors ^
i^
j^
k of the body-fixed frame.
To accomplish that, we proceed as follows.
The first rotation ½R3ðfÞ in Eqn (9.102) rotates the unit vectors ^
I^
J^
K of the inertial frame into the
unit vectors^
i1
^
j1
^
k1 of the intermediate x1y1z1 axes in Figure 9.23. Hence^
i1
^
j1
^
k1 are rotated into ^
I^
J^
K by
the inverse transformation given by
8

:
^
I
^
J
^
K
9


=


;
¼ ½R3ðfÞT
8





:
^
i1
^
j1
^
k1
9


=


;
¼
2
6
6
4
cos f sin f 0
sin f cos f 0
0 0 1
3
7
7
5
8





:
^
i1
^
j1
^
k1
9


=


;
(9.109)
512 CHAPTER 9 Rigid Body Dynamics
The second rotation [R1(q)] rotates^
i1
^
j1
^
k1 into the unit vectors^
i2
^
j2
^
k2 of the second intermediate frame
x2y2z2 in Figure 9.23. The inverse transformation rotates ^
i2
^
j2
^
k2 back into ^
i1
^
j1
^
k1:
8

:
^
i1
^
j1
^
k1
9


=


;
¼ ½R1ðqÞT
8





:
^
i2
^
j2
^
k2
9


=


;
¼
2
6
6
4
1 0 0
0 cos q sin q
0 sin q cos q
3
7
7
5
8





:
^
i2
^
j2
^
k2
9


=


;
(9.110)
Finally, the third rotation [R3(j)] rotates ^
i2
^
j2
^
k2 into ^
i^
j^
k, the target unit vectors of the body-fixed xyz
frame. ^
i2
^
j2
^
k2 are obtained from ^
i^
j^
k by the reverse rotation,
8

:
^
i2
^
j2
^
k2
9


=


;
¼ ½R3ðjÞT
8





:
^
i
^
j
^
k
9


=


;
¼
2
6
6
4
cos j sin j 0
sin j cos j 0
0 0 1
3
7
7
5
8





:
^
i
^
j
^
k
9


=


;
(9.111)
From Eqns (9.109), (9.110) and (9.111), we observe that
^
K
z}|{
9:109
¼ ^
k1
z}|{
9:110
¼ sinq^
j2 þ cosq^
k2
z}|{
9:111
¼ sinq

sinj^
i þ cosj^
j

þ cosq^
k
or
^
K ¼ sin q sin j^
i þ sin q cos j^
j þ cos q^
k (9.112)
Similarly, Eqns (9.110) and (9.111) imply that
^
i1 ¼ ^
i2 ¼ cos j^
i  sin j^
j (9.113)
Substituting Eqns (9.112) and (9.113) into Eqn (9.108) yields
u ¼ up

sin q sin j^
i þ sin q cos j^
j þ cos q^
k

þ un

cos j^
i  sin j^
j

þ us
^
k
or
u ¼

up sin q sin j þ un cos j

^
i þ

up sin q cos j  un sin j

^
j þ

us þ up cos q

^
k (9.114)
Comparing Eqns (9.107) and (9.114), we see that
ux ¼ up sin q sin j þ un cos j
uy ¼ up sin q cos j  un sin j
uz ¼ us þ up cos q
(9.115)
9.9 Euler angles 513
(Notice that the precession angle f does not appear.) We can solve these three equations to obtain the
Euler rates in terms of ux, uy, and uz:
up ¼ _
f ¼
1
sin q

ux sin j þ uy cos j

un ¼ _
q ¼ ux cos j  uy sin j
us ¼ _
j ¼ 
1
tan q

ux sin j þ uy cos j

þ uz
(9.116)
Observe that if ux, uy, and uz are given functions of time, found by solving Euler’s equations of motion
(Eqn (9.72)), then Eqns (9.116) are three coupled differential equations that may be solved to obtain
the three time-dependent Euler angles, namely
f ¼ fðtÞ q ¼ qðtÞ j ¼ jðtÞ
With this solution, the orientation of the xyz frame, and hence the body to which it is attached, is known
for any given time t. Note, however, that Eqns (9.116) “blow up” when q ¼ 0, that is, when the xy plane
is parallel to the XY plane.
EXAMPLE 9.18
At a given instant, the unit vectors of a body frame are
^
i ¼ 0:40825^
I  0:40825^
J þ 0:81649^
K
^
j ¼ 0:10102^
I  0:90914^
J  0:40405^
K
^
k ¼ 0:90726^
I þ 0:082479^
J  0:41240^
K
(a)
and the angular velocity is
u ¼ 3:1^
I þ 2:5^
J þ 1:7^
K ðrad=sÞ (b)
Calculate up, un, and us (the precession, nutation, and spin rates) at this instant.
Solution
We will ultimately use Eqn (9.116) to find up, un, and us. To do so we must first obtain the Euler angles f, q, and j
as well as the components of the angular velocity in the body frame.
The three rows of the direction cosine matrix [Q]Xx comprise the components of the unit vectors ^
i, ^
j, and ^
k,
respectively,
½QXx ¼
2
6
6
4
0:40825 0:40825 0:81649
0:10102 0:90914 0:40405
0:90726 0:082479 0:41240
3
7
7
5 (c)
Therefore, the components of the angular velocity in the body frame are
fugx ¼ ½QXx fugX ¼
2
6
6
4
0:40825 0:40825 0:81649
0:10102 0:90914 0:40405
0:90726 0:082479 0:41240
3
7
7
5
8





:
3:1
2:5
1:7
9


=


;
¼
8





:
0:89817
2:6466
3:3074
9


=


;
514 CHAPTER 9 Rigid Body Dynamics
or
ux ¼ 0:89817 rad=s uy ¼ 2:6466 rad=s uz ¼ 3:3074 rad=s (d)
To obtain the Euler angles f, q, and j from the direction cosine matrix in Eqn (c), we use Algorithm 4.3, as
illustrated in Example 9.17. That algorithm is implemented as the MATLAB function dcm_to_Euler.m
in Appendix D.20. Typing the following lines in the MATLAB Command Window:
 Q ¼ [ .40825 -.40825 .81649
-.10102 -.90914 -.40405
.90726 .082479 -.41240];
 [phi theta psi] ¼ dcm_to_euler(Q)
produces the following output:
phi ¼
95.1945
theta ¼
114.3557
psi ¼
116.3291
Substituting q ¼ 114.36 and j ¼ 116.33 together with the angular velocities of Eqn (d) into Eqn (9.116)
yields
up ¼
1
sin 114:36
½  0:89817$sin 116:33
þ ð2:6466Þ$cos 116:33
 ¼ 0:40492 rad=s
un ¼ 0:89817$cos 116:33
 ð2:6466Þ$sin 116:33
¼ 2:7704 rad=s
us ¼ 
1
tan 114:36
½  0:89817$sin 116:33
þ ð2:6466Þ$cos 116:33
 þ ð3:3074Þ ¼ 3:1404 rad=s
EXAMPLE 9.19
The mass moments of inertia of a body about the principal body frame axes with origin at the center of mass
G are
A ¼ 1000 kg$m2
B ¼ 2000 kg$m2
C ¼ 3000 kg$m2
(a)
The Euler angles in radians are given as functions of time in seconds as follows:
f ¼ 2te0:05t
q ¼ 0:02 þ 0:3 sin 0:25t
j ¼ 0:6t
(b)
At t ¼ 10 s, find
(a) The net moment about G and
(b) The components aX, aY, and aZ of the absolute angular acceleration in the inertial frame.
9.9 Euler angles 515
Solution
(a) We must use Euler’s equations (Eqns (9.72)) to calculate the net moment, which means we must first obtain
ux, uy, uz, _
ux , _
uy , and _
uz . Since we are given the Euler angles as a function of time, we can compute their time
derivatives and then use Eqn (9.115) to find the body frame angular velocity components and their derivatives.
Starting with Eqn (b), we get
up ¼
df
dt
¼
d
dt

2te0:05t

¼ 2e0:05t
 0:1te0:05t
_
up ¼
dup
dt
¼
d
dt

2e0:05t
 0:1e0:05t

¼ 0:2e0:05t
þ 0:005te0:05t
Proceeding to the remaining two Euler angles leads to
un ¼
dq
dt
¼
d
dt
ð0:02 þ 0:3 sin 0:25tÞ ¼ 0:075 cos 0:25t
_
un ¼
dun
dt
¼
d
dt
ð0:075 cos 0:25tÞ ¼ 0:01875 sin 0:25t
us ¼
dj
dt
¼
d
dt
ð0:6tÞ ¼ 0:6
_
us ¼
dus
dt
¼ 0
Evaluating all these quantities, including those in Eqn (b), at t ¼ 10 s yields
f ¼ 335:03 up ¼ 0:60653 rad=s _
up ¼ 0:09098 rad=s2
q ¼ 11:433 un ¼ 0:060086 rad=s _
un ¼ 0:011221 rad=s2
j ¼ 343:77 us ¼ 0:6 rad=s _
us ¼ 0
(c)
Equation (9.115) relates the Euler angle rates to the angular velocity components,
ux ¼ up sin q sin j þ un cos j
uy ¼ up sin q cos j  un sin j
uz ¼ us þ up cos q
(d)
Taking the time derivative of each of these equations in turn leads to the following three equations:
_
ux ¼ upun cos q sin j þ upus sin q cos j  unus sin j þ _
up sin q sin j þ _
un cos j
_
uy ¼ upun cos q cos j  up us sin q sin j  unus cos j þ _
up sin q cos j  _
un sin j
_
uz ¼ upun sin q þ _
up cos q þ _
us
(e)
Substituting the data in Eqn (c) into Eqns (d) and (e) yields
ux ¼ 0:091286 rad=s uy ¼ 0:098649 rad=s uz ¼ 1:1945 rad=s
_
ux ¼ 0:063435 rad=s2
_
uy ¼ 2:2346  105 rad=s2
_
uz ¼ 0:08195 rad=s2
(f)
With Eqns (a) and (f) we have everything we need for Euler’s equations, namely,
Mxnet
¼ A _
ux þ ðC  BÞuy uz
Mynet
¼ B _
uy þ ðA  CÞuz ux
Mznet
¼ C _
uz þ ðB  AÞux uy
516 CHAPTER 9 Rigid Body Dynamics
from which we find
Mxnet
¼ 181:27 N$m
Mynet
¼ 218:12 N$m
Mznet
¼ 254:86 N$m
(b) Since the comoving xyz frame is a body frame, rigidly attached to the solid, we know from Eqn (9.74) that
8

:
aX
aY
aZ
9


=


;
¼ ½QxX
8





:
_
ux
_
uy
_
uz
9


=


;
(g)
In other words, the absolute angular acceleration and the relative angular acceleration of the body are the
same. All we have to do is project the components of relative acceleration in Eqn (f) onto the axes of the
inertial frame. The required orthogonal transformation matrix is given in Eqn (9.105),
½QxX ¼
2
6
6
4
sin f cos q sin j þ cos f cos j sin f cos q cos g  cos f sin j sin f sin q
cos f cos q sin j þ sin f cos j cos f cos q cos j  sin f sin j cos f sin q
sin q sin j sin q cos j cos q
3
7
7
5
Upon substituting the numerical values of the Euler angles from Eqn (c), this becomes
½QxX ¼
2
6
6
4
0:75484 0:65055 0:083668
0:65356 0:73523 0:17970
0:055386 0:19033 0:98016
3
7
7
5
Substituting this and the relative angular velocity rates from Eqn (f) into Eqn (g) yields
8

:
aX
aY
aZ
9


=


;
¼
2
6
6
4
0:75484 0:65055 0:083668
0:65356 0:73523 0:17970
0:055386 0:19033 0:98016
3
7
7
5
8





:
0:063435
2:2345  105
0:08195
9


=


;
¼
8









:
0:054755
0:026716
0:083833
9




=




;

rad=s2

EXAMPLE 9.20
Figure 9.24 shows a rotating platform on which is mounted a rectangular parallelepiped shaft (with dimensions
b, h, and [) spinning about the inclined axis DE. If the mass of the shaft is m, and the angular velocities up and
us are constant, calculate the bearing forces at D and E as a function of f and j. Neglect gravity, since we are
interested only in the gyroscopic forces. (The small extensions shown at each end of the parallelepiped are just for
clarity; the distance between the bearings at D and E is [.)
Solution
The inertial XYZ frame is centered at O on the platform, and it is right handed ð^
I ^
J ¼ ^
KÞ. The origin of the right-
handed comoving body frame xyz is at the shaft’s center of mass G, and it is aligned with the symmetry axes of the
9.9 Euler angles 517
parallelepiped. The three Euler angles f, q, and j are shown in Figure 9.24. Since q is constant, the nutation rate is
zero (un ¼ 0). Thus, Eqn (9.115) reduces to
ux ¼ up sin q sin j uy ¼ up sin q cos j uz ¼ up cos q þ us (a)
Since up, us, and q are constant, it follows (recalling Eqn (9.106)) that
_
ux ¼ up us sin q cos j _
uy ¼ up us sin q sin j _
uz ¼ 0 (b)
The principal moments of inertia of the parallelepiped are (Figure 9.10(c))
A ¼ ‘x ¼
1
12
m

h2
þ ‘2

B ¼ ‘y ¼
1
12
m

b2
þ ‘2

C ¼ ‘z ¼
1
12
m

b2
þ h2

(c)
Figure 9.25 is a free-body diagram of the shaft. Let us assume that the bearings at D and E are such as to exert just
the six body frame components of force shown. Thus, D is a thrust bearing to which the axial torque TD is applied
from, say, a motor of some kind. At E, there is a simple journal bearing.
D
X
Y
Z
y
z
x
O
E
(Measured in XY plane)
(Measured in Zz plane)
Shaft
Platform
˙
p
˙ s
x
x
(Measured in x'x plane,
perpendicular to shaft)
xyz axes attached
to the shaft
l/2
G
l/2
b h
φ
φ
θ
ψ ψ
ω
ω
=
=
′
′
FIGURE 9.24
Spinning block mounted on a rotating platform.
G
x
y
z
b
h
TD
Dx
Dz
Dy
Ey
Ex
l/2
l/2
FIGURE 9.25
Free-body diagram of the block in Figure 9.24.
518 CHAPTER 9 Rigid Body Dynamics
9 rigid body dynamics.pdf
9 rigid body dynamics.pdf
9 rigid body dynamics.pdf
9 rigid body dynamics.pdf
9 rigid body dynamics.pdf
9 rigid body dynamics.pdf
9 rigid body dynamics.pdf
9 rigid body dynamics.pdf
9 rigid body dynamics.pdf
9 rigid body dynamics.pdf
9 rigid body dynamics.pdf
9 rigid body dynamics.pdf
9 rigid body dynamics.pdf
9 rigid body dynamics.pdf
9 rigid body dynamics.pdf
9 rigid body dynamics.pdf
9 rigid body dynamics.pdf
9 rigid body dynamics.pdf
9 rigid body dynamics.pdf
9 rigid body dynamics.pdf
9 rigid body dynamics.pdf
9 rigid body dynamics.pdf
9 rigid body dynamics.pdf
9 rigid body dynamics.pdf

More Related Content

Similar to 9 rigid body dynamics.pdf

6 maneuvres.pdf
6 maneuvres.pdf6 maneuvres.pdf
6 maneuvres.pdfa a
 
Ppt mech 5sem_dom
Ppt mech 5sem_domPpt mech 5sem_dom
Ppt mech 5sem_domvbrayka
 
Part IIWhen r= 0.7rPart II.docx
Part IIWhen r= 0.7rPart II.docxPart IIWhen r= 0.7rPart II.docx
Part IIWhen r= 0.7rPart II.docxherbertwilson5999
 
Amma kalviyagam-free-formula-hand-book
Amma kalviyagam-free-formula-hand-bookAmma kalviyagam-free-formula-hand-book
Amma kalviyagam-free-formula-hand-bookSenthilKumar Selvaraj
 
IEEE Paper .venkat (1).pdf
IEEE Paper .venkat (1).pdfIEEE Paper .venkat (1).pdf
IEEE Paper .venkat (1).pdfSheronThomas4
 
7 relative motion.pdf
7 relative motion.pdf7 relative motion.pdf
7 relative motion.pdfa a
 
analyzing system of motion of a particles
analyzing system of motion of a particlesanalyzing system of motion of a particles
analyzing system of motion of a particlesvikasaucea
 
Circular motion
Circular motionCircular motion
Circular motionArun Umrao
 
Principle of Circular Motion - Physics - An Introduction by Arun Umrao
Principle of Circular Motion - Physics - An Introduction by Arun UmraoPrinciple of Circular Motion - Physics - An Introduction by Arun Umrao
Principle of Circular Motion - Physics - An Introduction by Arun Umraossuserd6b1fd
 
Rotational Motion and Equilibrium
Rotational Motion and EquilibriumRotational Motion and Equilibrium
Rotational Motion and EquilibriumAJAL A J
 
Transfer function of Mechanical rotational system
Transfer function of Mechanical rotational system Transfer function of Mechanical rotational system
Transfer function of Mechanical rotational system KALPANA K
 

Similar to 9 rigid body dynamics.pdf (20)

6 maneuvres.pdf
6 maneuvres.pdf6 maneuvres.pdf
6 maneuvres.pdf
 
KOM-Unit-3.pptx
KOM-Unit-3.pptxKOM-Unit-3.pptx
KOM-Unit-3.pptx
 
Ppt mech 5sem_dom
Ppt mech 5sem_domPpt mech 5sem_dom
Ppt mech 5sem_dom
 
AP Physics C Rotational Motion
AP Physics C Rotational MotionAP Physics C Rotational Motion
AP Physics C Rotational Motion
 
Balancing
BalancingBalancing
Balancing
 
Part IIWhen r= 0.7rPart II.docx
Part IIWhen r= 0.7rPart II.docxPart IIWhen r= 0.7rPart II.docx
Part IIWhen r= 0.7rPart II.docx
 
15 bcm0081 assign 2
15 bcm0081 assign 215 bcm0081 assign 2
15 bcm0081 assign 2
 
Em notes
Em notesEm notes
Em notes
 
Physics 9
Physics 9Physics 9
Physics 9
 
Amma kalviyagam-free-formula-hand-book
Amma kalviyagam-free-formula-hand-bookAmma kalviyagam-free-formula-hand-book
Amma kalviyagam-free-formula-hand-book
 
Wang1998
Wang1998Wang1998
Wang1998
 
IEEE Paper .venkat (1).pdf
IEEE Paper .venkat (1).pdfIEEE Paper .venkat (1).pdf
IEEE Paper .venkat (1).pdf
 
12 rotational motion
12 rotational motion12 rotational motion
12 rotational motion
 
7 relative motion.pdf
7 relative motion.pdf7 relative motion.pdf
7 relative motion.pdf
 
analyzing system of motion of a particles
analyzing system of motion of a particlesanalyzing system of motion of a particles
analyzing system of motion of a particles
 
Circular motion
Circular motionCircular motion
Circular motion
 
Principle of Circular Motion - Physics - An Introduction by Arun Umrao
Principle of Circular Motion - Physics - An Introduction by Arun UmraoPrinciple of Circular Motion - Physics - An Introduction by Arun Umrao
Principle of Circular Motion - Physics - An Introduction by Arun Umrao
 
12 rotational motion 2
12 rotational motion 212 rotational motion 2
12 rotational motion 2
 
Rotational Motion and Equilibrium
Rotational Motion and EquilibriumRotational Motion and Equilibrium
Rotational Motion and Equilibrium
 
Transfer function of Mechanical rotational system
Transfer function of Mechanical rotational system Transfer function of Mechanical rotational system
Transfer function of Mechanical rotational system
 

More from a a

12 Perterbations.pdf
12 Perterbations.pdf12 Perterbations.pdf
12 Perterbations.pdfa a
 
4 3d orbits.pdf
4 3d orbits.pdf4 3d orbits.pdf
4 3d orbits.pdfa a
 
8 Interplanetary traj.pdf
8 Interplanetary traj.pdf8 Interplanetary traj.pdf
8 Interplanetary traj.pdfa a
 
3 position by time.pdf
3 position by time.pdf3 position by time.pdf
3 position by time.pdfa a
 
5 prelim determination.pdf
5 prelim determination.pdf5 prelim determination.pdf
5 prelim determination.pdfa a
 
11 rocket dyanmics.pdf
11 rocket dyanmics.pdf11 rocket dyanmics.pdf
11 rocket dyanmics.pdfa a
 
3 EDL.pdf
3 EDL.pdf3 EDL.pdf
3 EDL.pdfa a
 
1 Power Sys.pdf
1 Power Sys.pdf1 Power Sys.pdf
1 Power Sys.pdfa a
 
2 Thermal Control.pdf
2 Thermal Control.pdf2 Thermal Control.pdf
2 Thermal Control.pdfa a
 
59.pdf
59.pdf59.pdf
59.pdfa a
 
48.pdf
48.pdf48.pdf
48.pdfa a
 
50.pdf
50.pdf50.pdf
50.pdfa a
 
52.pdf
52.pdf52.pdf
52.pdfa a
 
56.pdf
56.pdf56.pdf
56.pdfa a
 
57.pdf
57.pdf57.pdf
57.pdfa a
 
55.pdf
55.pdf55.pdf
55.pdfa a
 
54.pdf
54.pdf54.pdf
54.pdfa a
 
51.pdf
51.pdf51.pdf
51.pdfa a
 
53.pdf
53.pdf53.pdf
53.pdfa a
 
47.pdf
47.pdf47.pdf
47.pdfa a
 

More from a a (20)

12 Perterbations.pdf
12 Perterbations.pdf12 Perterbations.pdf
12 Perterbations.pdf
 
4 3d orbits.pdf
4 3d orbits.pdf4 3d orbits.pdf
4 3d orbits.pdf
 
8 Interplanetary traj.pdf
8 Interplanetary traj.pdf8 Interplanetary traj.pdf
8 Interplanetary traj.pdf
 
3 position by time.pdf
3 position by time.pdf3 position by time.pdf
3 position by time.pdf
 
5 prelim determination.pdf
5 prelim determination.pdf5 prelim determination.pdf
5 prelim determination.pdf
 
11 rocket dyanmics.pdf
11 rocket dyanmics.pdf11 rocket dyanmics.pdf
11 rocket dyanmics.pdf
 
3 EDL.pdf
3 EDL.pdf3 EDL.pdf
3 EDL.pdf
 
1 Power Sys.pdf
1 Power Sys.pdf1 Power Sys.pdf
1 Power Sys.pdf
 
2 Thermal Control.pdf
2 Thermal Control.pdf2 Thermal Control.pdf
2 Thermal Control.pdf
 
59.pdf
59.pdf59.pdf
59.pdf
 
48.pdf
48.pdf48.pdf
48.pdf
 
50.pdf
50.pdf50.pdf
50.pdf
 
52.pdf
52.pdf52.pdf
52.pdf
 
56.pdf
56.pdf56.pdf
56.pdf
 
57.pdf
57.pdf57.pdf
57.pdf
 
55.pdf
55.pdf55.pdf
55.pdf
 
54.pdf
54.pdf54.pdf
54.pdf
 
51.pdf
51.pdf51.pdf
51.pdf
 
53.pdf
53.pdf53.pdf
53.pdf
 
47.pdf
47.pdf47.pdf
47.pdf
 

Recently uploaded

Islamabad Call Girls # 03091665556 # Call Girls in Islamabad | Islamabad Escorts
Islamabad Call Girls # 03091665556 # Call Girls in Islamabad | Islamabad EscortsIslamabad Call Girls # 03091665556 # Call Girls in Islamabad | Islamabad Escorts
Islamabad Call Girls # 03091665556 # Call Girls in Islamabad | Islamabad Escortswdefrd
 
exhuma plot and synopsis from the exhuma movie.pptx
exhuma plot and synopsis from the exhuma movie.pptxexhuma plot and synopsis from the exhuma movie.pptx
exhuma plot and synopsis from the exhuma movie.pptxKurikulumPenilaian
 
Lucknow 💋 Call Girl in Lucknow | Whatsapp No 8923113531 VIP Escorts Service A...
Lucknow 💋 Call Girl in Lucknow | Whatsapp No 8923113531 VIP Escorts Service A...Lucknow 💋 Call Girl in Lucknow | Whatsapp No 8923113531 VIP Escorts Service A...
Lucknow 💋 Call Girl in Lucknow | Whatsapp No 8923113531 VIP Escorts Service A...anilsa9823
 
Jeremy Casson - Top Tips for Pottery Wheel Throwing
Jeremy Casson - Top Tips for Pottery Wheel ThrowingJeremy Casson - Top Tips for Pottery Wheel Throwing
Jeremy Casson - Top Tips for Pottery Wheel ThrowingJeremy Casson
 
Aminabad @ Book Call Girls in Lucknow - 450+ Call Girl Cash Payment 🍵 8923113...
Aminabad @ Book Call Girls in Lucknow - 450+ Call Girl Cash Payment 🍵 8923113...Aminabad @ Book Call Girls in Lucknow - 450+ Call Girl Cash Payment 🍵 8923113...
Aminabad @ Book Call Girls in Lucknow - 450+ Call Girl Cash Payment 🍵 8923113...akbard9823
 
Deconstructing Gendered Language; Feminist World-Making 2024
Deconstructing Gendered Language; Feminist World-Making 2024Deconstructing Gendered Language; Feminist World-Making 2024
Deconstructing Gendered Language; Feminist World-Making 2024samlnance
 
Authentic # 00971556872006 # Hot Call Girls Service in Dubai By International...
Authentic # 00971556872006 # Hot Call Girls Service in Dubai By International...Authentic # 00971556872006 # Hot Call Girls Service in Dubai By International...
Authentic # 00971556872006 # Hot Call Girls Service in Dubai By International...home
 
RAK Call Girls Service # 971559085003 # Call Girl Service In RAK
RAK Call Girls Service # 971559085003 # Call Girl Service In RAKRAK Call Girls Service # 971559085003 # Call Girl Service In RAK
RAK Call Girls Service # 971559085003 # Call Girl Service In RAKedwardsara83
 
Call Girl Service In Dubai #$# O56521286O #$# Dubai Call Girls
Call Girl Service In Dubai #$# O56521286O #$# Dubai Call GirlsCall Girl Service In Dubai #$# O56521286O #$# Dubai Call Girls
Call Girl Service In Dubai #$# O56521286O #$# Dubai Call Girlsparisharma5056
 
this is a jarvis ppt for jarvis ai assistant lovers and this is for you
this is a jarvis ppt for jarvis ai assistant lovers and this is for youthis is a jarvis ppt for jarvis ai assistant lovers and this is for you
this is a jarvis ppt for jarvis ai assistant lovers and this is for youhigev50580
 
Lucknow 💋 Call Girls in Lucknow | Service-oriented sexy call girls 8923113531...
Lucknow 💋 Call Girls in Lucknow | Service-oriented sexy call girls 8923113531...Lucknow 💋 Call Girls in Lucknow | Service-oriented sexy call girls 8923113531...
Lucknow 💋 Call Girls in Lucknow | Service-oriented sexy call girls 8923113531...anilsa9823
 
Young⚡Call Girls in Lajpat Nagar Delhi >༒9667401043 Escort Service
Young⚡Call Girls in Lajpat Nagar Delhi >༒9667401043 Escort ServiceYoung⚡Call Girls in Lajpat Nagar Delhi >༒9667401043 Escort Service
Young⚡Call Girls in Lajpat Nagar Delhi >༒9667401043 Escort Servicesonnydelhi1992
 
Charbagh / best call girls in Lucknow - Book 🥤 8923113531 🪗 Call Girls Availa...
Charbagh / best call girls in Lucknow - Book 🥤 8923113531 🪗 Call Girls Availa...Charbagh / best call girls in Lucknow - Book 🥤 8923113531 🪗 Call Girls Availa...
Charbagh / best call girls in Lucknow - Book 🥤 8923113531 🪗 Call Girls Availa...gurkirankumar98700
 
Bridge Fight Board by Daniel Johnson dtjohnsonart.com
Bridge Fight Board by Daniel Johnson dtjohnsonart.comBridge Fight Board by Daniel Johnson dtjohnsonart.com
Bridge Fight Board by Daniel Johnson dtjohnsonart.comthephillipta
 
Call girls in Kanpur - 9761072362 with room service
Call girls in Kanpur - 9761072362 with room serviceCall girls in Kanpur - 9761072362 with room service
Call girls in Kanpur - 9761072362 with room servicediscovermytutordmt
 
Young⚡Call Girls in Uttam Nagar Delhi >༒9667401043 Escort Service
Young⚡Call Girls in Uttam Nagar Delhi >༒9667401043 Escort ServiceYoung⚡Call Girls in Uttam Nagar Delhi >༒9667401043 Escort Service
Young⚡Call Girls in Uttam Nagar Delhi >༒9667401043 Escort Servicesonnydelhi1992
 
Lucknow 💋 Call Girl in Lucknow 10k @ I'm VIP Independent Escorts Girls 892311...
Lucknow 💋 Call Girl in Lucknow 10k @ I'm VIP Independent Escorts Girls 892311...Lucknow 💋 Call Girl in Lucknow 10k @ I'm VIP Independent Escorts Girls 892311...
Lucknow 💋 Call Girl in Lucknow 10k @ I'm VIP Independent Escorts Girls 892311...anilsa9823
 
Lucknow 💋 Escort Service in Lucknow (Adult Only) 8923113531 Escort Service 2...
Lucknow 💋 Escort Service in Lucknow  (Adult Only) 8923113531 Escort Service 2...Lucknow 💋 Escort Service in Lucknow  (Adult Only) 8923113531 Escort Service 2...
Lucknow 💋 Escort Service in Lucknow (Adult Only) 8923113531 Escort Service 2...anilsa9823
 

Recently uploaded (20)

RAJKOT CALL GIRL 76313*77252 CALL GIRL IN RAJKOT
RAJKOT CALL GIRL 76313*77252 CALL GIRL IN RAJKOTRAJKOT CALL GIRL 76313*77252 CALL GIRL IN RAJKOT
RAJKOT CALL GIRL 76313*77252 CALL GIRL IN RAJKOT
 
Islamabad Call Girls # 03091665556 # Call Girls in Islamabad | Islamabad Escorts
Islamabad Call Girls # 03091665556 # Call Girls in Islamabad | Islamabad EscortsIslamabad Call Girls # 03091665556 # Call Girls in Islamabad | Islamabad Escorts
Islamabad Call Girls # 03091665556 # Call Girls in Islamabad | Islamabad Escorts
 
exhuma plot and synopsis from the exhuma movie.pptx
exhuma plot and synopsis from the exhuma movie.pptxexhuma plot and synopsis from the exhuma movie.pptx
exhuma plot and synopsis from the exhuma movie.pptx
 
Lucknow 💋 Call Girl in Lucknow | Whatsapp No 8923113531 VIP Escorts Service A...
Lucknow 💋 Call Girl in Lucknow | Whatsapp No 8923113531 VIP Escorts Service A...Lucknow 💋 Call Girl in Lucknow | Whatsapp No 8923113531 VIP Escorts Service A...
Lucknow 💋 Call Girl in Lucknow | Whatsapp No 8923113531 VIP Escorts Service A...
 
Jeremy Casson - Top Tips for Pottery Wheel Throwing
Jeremy Casson - Top Tips for Pottery Wheel ThrowingJeremy Casson - Top Tips for Pottery Wheel Throwing
Jeremy Casson - Top Tips for Pottery Wheel Throwing
 
Aminabad @ Book Call Girls in Lucknow - 450+ Call Girl Cash Payment 🍵 8923113...
Aminabad @ Book Call Girls in Lucknow - 450+ Call Girl Cash Payment 🍵 8923113...Aminabad @ Book Call Girls in Lucknow - 450+ Call Girl Cash Payment 🍵 8923113...
Aminabad @ Book Call Girls in Lucknow - 450+ Call Girl Cash Payment 🍵 8923113...
 
Deconstructing Gendered Language; Feminist World-Making 2024
Deconstructing Gendered Language; Feminist World-Making 2024Deconstructing Gendered Language; Feminist World-Making 2024
Deconstructing Gendered Language; Feminist World-Making 2024
 
Authentic # 00971556872006 # Hot Call Girls Service in Dubai By International...
Authentic # 00971556872006 # Hot Call Girls Service in Dubai By International...Authentic # 00971556872006 # Hot Call Girls Service in Dubai By International...
Authentic # 00971556872006 # Hot Call Girls Service in Dubai By International...
 
RAK Call Girls Service # 971559085003 # Call Girl Service In RAK
RAK Call Girls Service # 971559085003 # Call Girl Service In RAKRAK Call Girls Service # 971559085003 # Call Girl Service In RAK
RAK Call Girls Service # 971559085003 # Call Girl Service In RAK
 
Call Girl Service In Dubai #$# O56521286O #$# Dubai Call Girls
Call Girl Service In Dubai #$# O56521286O #$# Dubai Call GirlsCall Girl Service In Dubai #$# O56521286O #$# Dubai Call Girls
Call Girl Service In Dubai #$# O56521286O #$# Dubai Call Girls
 
Dxb Call Girls # +971529501107 # Call Girls In Dxb Dubai || (UAE)
Dxb Call Girls # +971529501107 # Call Girls In Dxb Dubai || (UAE)Dxb Call Girls # +971529501107 # Call Girls In Dxb Dubai || (UAE)
Dxb Call Girls # +971529501107 # Call Girls In Dxb Dubai || (UAE)
 
this is a jarvis ppt for jarvis ai assistant lovers and this is for you
this is a jarvis ppt for jarvis ai assistant lovers and this is for youthis is a jarvis ppt for jarvis ai assistant lovers and this is for you
this is a jarvis ppt for jarvis ai assistant lovers and this is for you
 
Lucknow 💋 Call Girls in Lucknow | Service-oriented sexy call girls 8923113531...
Lucknow 💋 Call Girls in Lucknow | Service-oriented sexy call girls 8923113531...Lucknow 💋 Call Girls in Lucknow | Service-oriented sexy call girls 8923113531...
Lucknow 💋 Call Girls in Lucknow | Service-oriented sexy call girls 8923113531...
 
Young⚡Call Girls in Lajpat Nagar Delhi >༒9667401043 Escort Service
Young⚡Call Girls in Lajpat Nagar Delhi >༒9667401043 Escort ServiceYoung⚡Call Girls in Lajpat Nagar Delhi >༒9667401043 Escort Service
Young⚡Call Girls in Lajpat Nagar Delhi >༒9667401043 Escort Service
 
Charbagh / best call girls in Lucknow - Book 🥤 8923113531 🪗 Call Girls Availa...
Charbagh / best call girls in Lucknow - Book 🥤 8923113531 🪗 Call Girls Availa...Charbagh / best call girls in Lucknow - Book 🥤 8923113531 🪗 Call Girls Availa...
Charbagh / best call girls in Lucknow - Book 🥤 8923113531 🪗 Call Girls Availa...
 
Bridge Fight Board by Daniel Johnson dtjohnsonart.com
Bridge Fight Board by Daniel Johnson dtjohnsonart.comBridge Fight Board by Daniel Johnson dtjohnsonart.com
Bridge Fight Board by Daniel Johnson dtjohnsonart.com
 
Call girls in Kanpur - 9761072362 with room service
Call girls in Kanpur - 9761072362 with room serviceCall girls in Kanpur - 9761072362 with room service
Call girls in Kanpur - 9761072362 with room service
 
Young⚡Call Girls in Uttam Nagar Delhi >༒9667401043 Escort Service
Young⚡Call Girls in Uttam Nagar Delhi >༒9667401043 Escort ServiceYoung⚡Call Girls in Uttam Nagar Delhi >༒9667401043 Escort Service
Young⚡Call Girls in Uttam Nagar Delhi >༒9667401043 Escort Service
 
Lucknow 💋 Call Girl in Lucknow 10k @ I'm VIP Independent Escorts Girls 892311...
Lucknow 💋 Call Girl in Lucknow 10k @ I'm VIP Independent Escorts Girls 892311...Lucknow 💋 Call Girl in Lucknow 10k @ I'm VIP Independent Escorts Girls 892311...
Lucknow 💋 Call Girl in Lucknow 10k @ I'm VIP Independent Escorts Girls 892311...
 
Lucknow 💋 Escort Service in Lucknow (Adult Only) 8923113531 Escort Service 2...
Lucknow 💋 Escort Service in Lucknow  (Adult Only) 8923113531 Escort Service 2...Lucknow 💋 Escort Service in Lucknow  (Adult Only) 8923113531 Escort Service 2...
Lucknow 💋 Escort Service in Lucknow (Adult Only) 8923113531 Escort Service 2...
 

9 rigid body dynamics.pdf

  • 1. Rigid Body Dynamics 9 CHAPTER OUTLINE 9.1 Introduction ..................................................................................................................................459 9.2 Kinematics....................................................................................................................................460 9.3 Equations of translational motion ...................................................................................................469 9.4 Equations of rotational motion........................................................................................................470 9.5 Moments of inertia ........................................................................................................................475 Parallel axis theorem ...........................................................................................................490 9.6 Euler’s equations...........................................................................................................................496 9.7 Kinetic energy...............................................................................................................................502 9.8 The spinning top ...........................................................................................................................504 9.9 Euler angles..................................................................................................................................510 9.10 Yaw, pitch, and roll angles ..........................................................................................................520 9.11 Quaternions ................................................................................................................................523 Problems.............................................................................................................................................532 Section 9.2 .........................................................................................................................................532 Section 9.5 .........................................................................................................................................535 Section 9.7 .........................................................................................................................................540 Section 9.8 .........................................................................................................................................540 Section 9.9 .........................................................................................................................................542 9.1 Introduction Just as Chapter 1 provides a foundation for the dof the equations of orbital mechanics, this chapter serves as a basis for developing the equations of satellite attitude dynamics. Chapter 1 deals with particles, whereas here we are concerned with rigid bodies. Those familiar with rigid body dynamics can move on to the next chapter, perhaps returning from time to time to review concepts. The kinematics of rigid bodies is presented first. The subject depends on a theorem of the French mathematician Michel Chasles (1793–1880). Chasles’ theorem states that the motion of a rigid body can be described by the displacement of any point of the body (the base point) plus a rotation about a unique axis through that point. The magnitude of the rotation does not depend on the base point. Thus, at any instant, a rigid body in a general state of motion has an angular velocity vector whose direction is that of the instantaneous axis of rotation. Describing the rotational component of the motion a rigid body in three dimensions requires taking advantage of the vector nature of angular velocity and knowing how to take the time derivative of moving vectors, which is explained in Chapter 1. Several examples illustrate how this is done. CHAPTER Orbital Mechanics for Engineering Students. http://dx.doi.org/10.1016/B978-0-08-097747-8.00009-8 Copyright Ó 2014 Elsevier Ltd. All rights reserved. 459
  • 2. We then move on to study the interaction between the motion of a rigid body and the forces acting on it. Describing the translational component of the motion requires simply concentrating all of the mass at a point, thecenter of mass, andapplying themethodsof particle mechanics to determineits motion.Indeed, our study of the two-body problem up to this point has focused on the motion of their centers of mass without regard to the rotational aspect. Analyzing the rotational dynamics requires computing the body’s angular momentum, and that in turn requires accounting for how the mass is distributed throughout the body. The mass distribution is described by the six components of the moment of inertia tensor. Writing the equations of rotational motion relative to coordinate axes embedded in the rigid body and aligned with the principal axes of inertia yields the nonlinear Euler equations of motion, which are applied to a study of the dynamics of a spinning top (or one-axis gyro). The expression for the kinetic energy of a rigid body is derived because it will be needed in the following chapter. The chapter next describes the two sets of three angles commonly employed to specify the orien- tation of a body in three-dimensional space. One of these comprises the Euler angles, which are the same as the right ascension of the node (U), argument of periapsis (u), and inclination (i) introduced in Chapter 4 to orient orbits in space. The other set comprises the yaw, pitch, and roll angles, which are suitable for describing the orientation of an airplane. Both the Euler angles and yaw, pitch, and roll angles will be employed in Chapter 10. The chapter concludes with a brief discussion of quaternions and an example of how they are used to describe the evolution of the attitude of a rigid body. 9.2 Kinematics Figure 9.1 shows a moving rigid body and its instantaneous axis of rotation, which defines the direction of the absolute angular velocity vector u. The XYZ axes are a fixed, inertial frame of reference. The position vectors RA and RB of two points on the rigid body are measured in the inertial frame. The vector RB/A drawn from point A to point B is the position vector of B relative to A. Since the body is rigid, RB/A has a constant magnitude even though its direction is continuously changing. Clearly, RB ¼ RA þ RB=A Differentiating this equation through with respect to time, we get _ RB ¼ _ RA þ dRB=A dt (9.1) _ RA and _ RB are the absolute velocities vA and vB of points A and B. Because the magnitude of RB/A does not change, its time derivative is given by Eqn (1.61), that is, dRB=A dt ¼ u RB=A Thus, Eqn (9.1) becomes vB ¼ vA þ u RB=A (9.2) Taking the time derivative of Eqn (9.1) yields € RB ¼ € RA þ d2 RB=A dt2 (9.3) 460 CHAPTER 9 Rigid Body Dynamics
  • 3. € RA and € RB are the absolute accelerations aA and aB of the two points of the rigid body, while from Eqn (1.62) we have d2 RB=A dt2 ¼ a RB=A þ u u RB=A in which a is the angular acceleration, a ¼ du=dt. Therefore, Eqn (9.3) can be written aB ¼ aA þ a RB=A þ u u RB=A (9.4) Equations (9.2) and (9.4) are the relative velocity and acceleration formulas. Note that all quantities in these expressions are measured in the same inertial frame of reference. When the rigid body under consideration is connected to and moving relative to another rigid body, computation of its inertial angular velocity u and angular acceleration a must be done with care. The key is to remember that angular velocity is a vector. It may be found as the vector sum of a sequence of angular velocities, each measured relative to another, starting with one measured relative to an ab- solute frame, as illustrated in Figure 9.2. In that case, the absolute angular velocity u of body 4 is u ¼ u1 þ u2=1 þ u3=2 þ u4=3 (9.5) Each of these angular velocities is resolved into components along the axes of the moving frame of reference xyz as shown in Figure 9.2, so that the vector sum may be written as u ¼ ux ^ i þ uy ^ j þ uz ^ k (9.6) The moving frame is chosen for convenience of the analysis, and its own inertial angular velocity is denoted as U, as discussed in Section 1.6. According to Eqn (1.65), the absolute angular acceleration a ¼ du=dt is obtained from Eqn (9.6) by means of the following calculation: a ¼ du dt rel þ U u (9.7) X Y Z A RB RA B RB/A FIGURE 9.1 A rigid body and its instantaneous axis of rotation. 9.2 Kinematics 461
  • 4. where du dt rel ¼ _ ux ^ i þ _ uy ^ j þ _ uz ^ k (9.8) and _ ux ¼ dux=dt, etc. Being able to express the absolute angular velocity vector in an appropriately chosen moving reference frame, as in Eqn (9.6), is crucial to the analysis of rigid body motion. Once we have the components of u, we simply differentiate them with respect to time to arrive at Eqn (9.8). Observe that the absolute angular acceleration a and du=dtÞrel, the angular acceleration relative to the moving frame, are the same if and only if U ¼ u. That occurs if the moving reference is a body-fixed frame, that is, a set of xyz axes imbedded in the rigid body itself. EXAMPLE 9.1 The airplane in Figure 9.3 flies at a constant speed v while simultaneously undergoing a constant yaw rate uyaw about a vertical axis and describing a circular loop in the vertical plane with a radius r. The constant propeller spin rate is uspin relative to the airframe. Find the velocity and acceleration of the tip P of the propeller relative to the hub H, when P is directly above H. The propeller radius is l. Solution The xyz axes are rigidly attached to the airplane. The x-axis is aligned with the propeller’s spin axis. The y axis is vertical, and the z axis is in the spanwise direction, so that xyz forms a right-handed triad. Although the xyz frame is not inertial, we can imagine it to instantaneously coincide with an inertial frame. The absolute angular velocity of the airplane has two components, the yaw rate and the counterclockwise pitch angular velocity v/r of its rotation in the circular loop, uairplane ¼ uyaw ^ j þ upitch ^ k ¼ uyaw ^ j þ v r ^ k FIGURE 9.2 Angular velocity is the vector sum of the relative angular velocities starting with u1, measured relative to the inertial frame. 462 CHAPTER 9 Rigid Body Dynamics
  • 5. The angular velocity of the body-fixed moving frame is that of the airplane, U ¼ uairplane, so that U ¼ uyaw ^ j þ v r ^ k (a) The absolute angular velocity of the propeller is that of the airplane plus the angular velocity of the propeller relative to the airplane uprop ¼ uairplane þ uspin ^ i ¼ U þ uspin ^ i which means uprop ¼ uspin ^ i þ uyaw ^ j þ v r ^ k (b) From Eqn (9.2), the velocity of point P on the propeller relative to H on the hub, vP/H, is given by vP=H ¼ vP vH ¼ uprop rP=H where rP/H is the position vector of P relative to H at this instant, rP=H ¼ l^ j (c) Thus, using Eqns (b) and (c), vP=H ¼ uspin ^ i þ uyaw ^ j þ v r ^ k l^ j from which vP=H ¼ v r l^ i þ uspinl^ k The absolute angular acceleration of the propeller is found by substituting Eqns (a) and (b) into Eqn (9.7), aprop ¼ duprop dt rel þ U uprop ¼ duspin dt ^ i þ duyaw dt ^ j þ dðv=rÞ dt ^ k þ uyaw ^ j þ v r ^ k uspin ^ i þ uyaw ^ j þ v r ^ k Since uspin, uyaw, v, and r are all constant, this reduces to aprop ¼ uyaw ^ j þ v r ^ k uspin ^ i þ uyaw ^ j þ v r ^ k FIGURE 9.3 Airplane with an attached xyz body frame. 9.2 Kinematics 463
  • 6. Carrying out the cross product yields aprop ¼ v r uspin ^ j uyawuspin ^ k (d) From Eqn (9.4), the acceleration of P relative to H, aP/H, is given by aP=H ¼ aP aH ¼ aprop rP=H þ uprop uprop rP=H Substituting Eqns (b), (c) and (d) into this expression yields aP=H ¼ v r uspin ^ j uyawuspin ^ k l^ j þ uspin ^ i þ uyaw ^ j þ v r ^ k uspin ^ i þ uyaw ^ j þ v r ^ k l^ j From this we find that aP=H ¼ uyawuspinl^ i þ uspin ^ i þ uyaw ^ j þ v r ^ k v r l^ i þ uspinl^ k ¼ uyawuspinl^ i þ uyawuspinl^ i v2 r2 þ u2 spin l^ j þ uyaw v r l^ k so that finally, aP=H ¼ 2uyawuspinl^ i v2 r2 þ u2 spin l^ j þ uyaw v r l^ k EXAMPLE 9.2 The satellite in Figure 9.4 is rotating about the z axis at a constant rate N. The xyz axes are attached to the spacecraft, and the z-axis has a fixed orientation in inertial space. The solar panels rotate at a constant rate _ q in the direction shown. Relative to point O, which lies at the center of the spacecraft and on the centerline of the panels, calculate for point A on the panel (a) Its absolute velocity and (b) Its absolute acceleration. FIGURE 9.4 Rotating solar panel on a rotating satellite. 464 CHAPTER 9 Rigid Body Dynamics
  • 7. Solution (a) Since the moving xyz frame is attached to the body of the spacecraft, its angular velocity is U ¼ N^ k (a) The absolute angular velocity of the panel is the absolute angular velocity of the spacecraft plus the angular velocity of the panel relative to the spacecraft, upanel ¼ _ q^ j þ N^ k (b) The position vector of A relative to O is rA=O ¼ w 2 sin q^ i þ d^ j þ w 2 cos q^ k (c) According to Eqn (9.2), the velocity of A relative to O is vA=O ¼ vA vO ¼ upanel rA=O ¼ ^ i ^ j ^ k 0 _ q N w 2 sin q d w 2 cos q from which we get vA=O ¼ w 2 _ q cos q þ Nd ^ i w 2 N sin q^ j w 2 _ q sin q^ k (b) The absolute angular acceleration of the panel is found by substituting Eqns (a) and (b) into Eqn (9.7), apanel ¼ dupanel dt rel þ U upanel ¼ d _ q dt ^ j þ dN dt ^ k # þ N^ k _ q^ j þ N^ k Since N and _ q are constants, this reduces to apanel ¼ _ qN^ i (d) To find the acceleration of A relative to O, we substitute Eqns (b)e(d) into Eqn (9.4), aA=O ¼ aA aO ¼ apanel rA=O þ upanel upanel rA=O ¼ ^ i ^ j ^ k _ qN 0 0 w 2 sin q d w 2 cos q þ _ q^ j þ N^ k ^ i ^ j ^ k 0 _ q N w 2 sin q d w 2 cos q ¼ w 2 N _ q cos q^ j þ N _ qd^ k þ ^ i ^ j ^ k 0 _ q N w 2 _ q cos q Nd N w 2 sin q w 2 _ q sin q 9.2 Kinematics 465
  • 8. which leads to aA=O ¼ w 2 N2 þ _ q 2 sin q^ i N Nd þ w _ q cos q ^ j w 2 _ q 2 cos q^ k EXAMPLE 9.3 The gyro rotor illustrated in Figure 9.5 has a constant spin rate uspin around axis bea in the direction shown. The XYZ axes are fixed. The xyz axes are attached to the gimbal ring, whose angle q with the vertical is increasing at a constant rate _ q in the direction shown. The assembly is forced to precess at a constant rate N around the vertical, For the rotor in the position shown, calculate (a) The absolute angular velocity and (b) The absolute angular acceleration. Express the results in both the fixed XYZ frame and the moving xyz frame. Solution (a) We will need the instantaneous relationship between the unit vectors of the inertial XYZ axes and the comoving xyz frame, which on inspecting Figure 9.6 can be seen to be ^ I ¼ cos q^ j þ sin q^ k ^ J ¼ ^ i ^ K ¼ sin q^ j þ cos q^ k (a) so that the matrix of the transformation from xyz to XYZ is (Section 4.5) ½QxX ¼ 2 6 6 4 0 cos q sin q 1 0 0 0 sin q cos q 3 7 7 5 (b) The absolute angular velocity of the gimbal ring is that of the base plus the angular velocity of the gimbal relative to the base given as ugimbal ¼ N^ K þ _ q^ i ¼ N sin q^ j þ cos q^ k þ _ q^ i ¼ _ q^ i þ N sin q^ j þ N cos q^ k (c) where we made use of Eqn (a) above. Since the moving xyz frame is attached to the gimbal, U ¼ ugimbal, so that U ¼ _ q^ i þ N sin q^ j þ N cos q^ k (d) The absolute angular velocity of the rotor is its spin relative to the gimbal, plus the angular velocity of the gimbal, urotor ¼ ugimbal þ uspin ^ k (e) From Eqn (c), it follows that urotor ¼ _ q^ i þ N sin q^ j þ N cos q þ uspin ^ k (f) 466 CHAPTER 9 Rigid Body Dynamics
  • 9. θ θ Z FIGURE 9.6 Orientation of the fixed XZ axes relative to the rotating xz axes. FIGURE 9.5 Rotating, precessing, nutating gyro. 9.2 Kinematics 467
  • 10. Because^ i, ^ j and ^ k move with the gimbal, expression (f) is valid for any time, not just the instant shown in Figure 9.5. Alternatively, applying the vector transformation furotorgXYZ ¼ ½QxX furotorgxyz (g) we obtain the angular velocity of the rotor in the inertial frame, but only at the instant shown in the figure, that is, when the x-axis aligns with the Y-axis. 8 : uX uY uZ 9 = ; ¼ 2 6 6 6 4 0 cos q sin q 1 0 0 0 sin q cos q 3 7 7 7 5 8 : _ q N sin q N cos q þ uspin 9 = ; ¼ 8 : N sin q cos q þ N sin q cos q þ uspin sin q _ q N sin2 q þ N cos2q þ uspin cos q 9 = ; or urotor ¼ uspin sin q^ I þ _ q^ J þ N þ uspin cos q ^ K (h) (b) The angular acceleration of the rotor can be found by substituting Eqns (d) and (f) into Eqn (9.7), recalling that N, _ q, and uspin are independent of time: arotor ¼ durotor dt rel þ U urotor ¼ d _ q dt ^ i þ dðN sin qÞ dt ^ j þ d N cos q þ uspin dt ^ k # þ ^ i ^ j ^ k _ q N sin q N cos q _ q N sin q N cos q þ uspin ¼ N _ q cos q^ j N _ q sin q^ k þ h ^ i Nuspin sin q ^ j uspin _ q þ ^ kð0Þ i Upon collecting terms, we get arotor ¼ Nuspin sin q^ i þ _ q N cos q uspin ^ j N _ q sin q^ k (i) This expression, like Eqn (f), is valid at any time. The components of arotor along the XYZ axes are found in the same way as for urotor, farotorgXYZ ¼ ½QxX farotorgxyz which means 8 : aX aY aZ 9 = ; ¼ 2 6 6 6 4 0 cos q sin q 1 0 0 0 sin q cos q 3 7 7 7 5 8 : Nuspin sin q _ q N cos q uspin N _ q sin q 9 = ; ¼ 8 : N _ qcos2q þ _ quspin cos q N _ q sin2 q Nuspin sin q N _ q sin q cos q _ quspin sin q N _ q sin q cos q 9 = ; 468 CHAPTER 9 Rigid Body Dynamics
  • 11. or arotor ¼ _ q uspin cos q N ^ I þ Nuspin sin q^ J _ quspin sin q^ K (j) Note carefully that Eqn (j) is not simply the time derivative of Eqn (h). Equations (h) and (j) are valid only at the instant that the xyz and XYZ axes have the alignments shown in Figure 9.5. 9.3 Equations of translational motion Figure 9.7 again shows an arbitrary, continuous, three-dimensional body of mass m. “Continuous” means that as we zoom in on a point it remains surrounded by a continuous distribution of matter having the infinitesimal mass dm in the limit. The point never ends up in a void. In particular, we ignore the actual atomic and molecular microstructure in favor of this continuum hypothesis, as it is called. Molecular microstructure does bear upon the overall dynamics of a finite body. We will use G to denote the center of mass. The position vectors of points relative to the origin of the inertial frame will be designated by capital letters. Thus, the position of the center of mass is RG defined as mRG ¼ Z m Rdm (9.9) R is the position of a mass element dm within the continuum. Each element of mass is acted upon by a net external force dFnet and a net internal force dfnet. The external force comes from direct contact with other objects and from action at a distance, such as gravitational attraction. The internal forces are those exerted from within the body by neighboring particles. These are the forces that hold the body together. For each mass element, Newton’s second law, Eqn (1.47), is written as dFnet þ dfnet ¼ dm€ R (9.10) Writing this equation for the infinite number of mass elements of which the body is composed, and then summing them all together leads, to the integral Z dFnet þ Z dfnet ¼ Z m € Rdm FIGURE 9.7 Forces on the mass element dm of a continuous medium. 9.3 Equations of translational motion 469
  • 12. Because the internal forces occur in action–reaction pairs, R dfnet ¼ 0. (External forces on the body are those without an internal reactant; the reactant lies outside the body and, hence, is outside our pur- view.) Thus, Fnet ¼ Z m € Rdm (9.11) where Fnet is the resultant external force on the body, Fnet ¼ R dFnet. From Eqn (9.9), Z m € Rdm ¼ m€ RG where € RG ¼ aG, the absolute acceleration of the center of mass. Therefore, Eqn (9.11) can be written as Fnet ¼ m€ RG (9.12) We are therefore reminded that the motion of the center of mass of a body is determined solely by the resultant of the external forces acting on it. So far, our study of orbiting bodies has focused exclusively on the motion of their centers of mass. In this chapter, we will turn our attention to rotational motion around the center of mass. To simplify things, we will ultimately assume that the body is not only continuous but that it is also rigid. This means all points of the body remain at a fixed distance from each other and there is no flexing, bending, or twisting deformation. 9.4 Equations of rotational motion Our development of the rotational dynamics equations does not require at the outset that the body under consideration be rigid. It may be a solid, fluid, or gas. Point P in Figure 9.8 is arbitrary; it need not be fixed in space nor attached to a point on the body. Then, the moment about P of the forces on mass element dm (cf. Figure 9.7) is dMP ¼ r dFnet þ r dfnet where r is the position vector of the mass element dm relative to the point P. Writing the right-hand side as r ðdFnet þ dfnetÞ, substituting Eqn (9.10), and integrating over all the mass elements of the body yields MPnet ¼ Z m r € Rdm (9.13) where € R is the absolute acceleration of dm relative to the inertial frame and MPnet ¼ Z r dFnet þ Z r dfnet 470 CHAPTER 9 Rigid Body Dynamics
  • 13. But R r dfnet ¼ 0 because the internal forces occur in action–reaction pairs. Thus, MPnet ¼ Z r dFnet which means the net moment includes only the moment of all of the external forces on the body. From the product rule of calculus, we know that dðr _ RÞ=dt ¼ r € R þ _ r _ R, so that the inte- grand in Eqn (9.13) may be written as r € R ¼ d dt r _ R _ r _ R (9.14) Furthermore, Figure 9.8 shows that r ¼ R RP, where RP is the absolute position vector of P. It follows that _ r _ R ¼ _ R _ RP _ R ¼ _ RP _ R (9.15) Substituting Eqn (9.15) into Eqn (9.14), then moving that result into Eqn (9.13), yields MPnet ¼ d dt Z m r _ Rdm þ _ RP Z m _ Rdm (9.16) Now, r _ Rdm is the moment of the absolute linear momentum of mass element dm about P. The moment of momentum, or angular momentum, of the entire body is the integral of this cross- product over all of its mass elements. That is, the absolute angular momentum of the body relative to point P is HP ¼ Z m r _ Rdm (9.17) FIGURE 9.8 Position vectors of a mass element in a continuum from several key reference points. 9.4 Equations of rotational motion 471
  • 14. Observing from Figure 9.8 that r ¼ rG/P þ r, we can write Eqn (9.17) as HP ¼ Z m rG=P þ r _ Rdm ¼ rG=P Z m _ Rdm þ Z m r _ Rdm (9.18) The last term is the absolute angular momentum relative to the center of mass G, HG ¼ Z r _ Rdm (9.19) Furthermore, by the definition of center of mass, Eqn (9.9), Z m _ Rdm ¼ m _ RG (9.20) Equations (9.19) and (9.20) allow us to write Eqn (9.18) as HP ¼ HG þ rG=P mvG (9.21) This useful relationship shows how to obtain the absolute angular momentum about any point P once HG is known. For calculating the angular momentum about the center of mass, Eqn (9.19) can be cast in a much more useful form by making the substitution (cf. Figure 9.8) R ¼ RG þ r, so that HG ¼ Z m r _ RG þ _ r dm ¼ Z m r _ RGdm þ Z m r _ rdm In the two integrals on the right, the variable is r. _ RG is fixed and can therefore be factored out of the first integral to obtain HG ¼ 0 @ Z m rdm 1 A _ RG þ Z m r _ rdm By definition of the center of mass, R m rdm ¼ 0 (the position vector of the center of mass relative to itself is zero), which means HG ¼ Z m r _ rdm (9.22) Since r and _ r are the position and velocity relative to the center of mass G, R m r _ rdm is the total moment about the center of mass of the linear momentum relative to the center of mass, HGrel . In other words, HG ¼ HGrel (9.23) This is a rather surprising fact, hidden in Eqn (9.19), and is true in general for no other point of the body. 472 CHAPTER 9 Rigid Body Dynamics
  • 15. Another useful angular momentum formula, similar to Eqn (9.21), may be found by substituting R ¼ RP þ r into Eqn (9.17), HP ¼ Z m r _ RP þ _ r dm ¼ 0 @ Z m rdm 1 A _ RP þ Z m r _ rdm (9.24) The term on the far right is the net moment of relative linear momentum about P, HPrel ¼ Z m r _ rdm (9.25) Also, R m rdm ¼ mrG=P, where rG/P is the position of the center of mass relative to P. Thus, Eqn (9.24) can be written as HP ¼ HPrel þ rG=P mvP (9.26) Finally, substituting this into Eqn (9.21), solving for HPrel , and noting that vG vP ¼ vG/P yields HPrel ¼ HG þ rG=P mvG=P (9.27) This expression is useful when the absolute velocity vG of the center of mass, which is required in Eqn (9.21), is not available. So far, we have written down some formulas for calculating the angular momentum about an arbitrary point in space and about the center of mass of the body itself. Let us now return to the problem of relating angular momentum to the applied torque. Substituting Eqns (9.17) and (9.20) into Eqn (9.16), we obtain MPnet ¼ _ HP þ _ RP m _ RG Thus, for an arbitrary point P, MPnet ¼ _ HP þ vP mvG (9.28) where vP and vG are the absolute velocities of points P and G, respectively. This expression is applicable to two important special cases. If the point P is at rest in inertial space ðvP ¼ 0Þ, then Eqn (9.28) reduces to MPnet ¼ _ HP (9.29) This equation holds as well if vP and vG are parallel. If P is the point of contact of a wheel rolling while slipping in the plane. Note that the validity of Eqn (9.29) depends neither on the body’s being rigid nor on its being in pure rotation about P. If point P is chosen to be the center of mass, then, since vG vG ¼ 0, Eqn (9.28) becomes MGnet ¼ _ HG (9.30) This equation is valid for any state of motion. 9.4 Equations of rotational motion 473
  • 16. If Eqn (9.30) is integrated over a time interval, then we obtain the angular impulse–momentum principle, Zt2 t1 MGnet dt ¼ HG2 HG1 (9.31) A similar expression follows from Eqn (9.29). R Mdt is the angular impulse. If the net angular impulse is zero, then DH ¼ 0, which is a statement of the conservation of angular momentum. Keep in mind that the angular impulse–momentum principle is not valid for just any reference point. Additional versions of Eqns (9.29) and (9.30) can be obtained which may prove useful in special circumstances. For example, substituting the expression for HP (Eqn (9.21)) into Eqn (9.28) yields MPnet ¼ _ HG þ d dt rG=P mvG þ vP mvG ¼ _ HG þ d dt ½ðrG rPÞ mvG þ vP mvG ¼ _ HG þ ðvG vPÞ mvG þ rG=P maG þ vP mvG or, finally, MPnet ¼ _ HG þ rG=P maG (9.32) This expression is useful when it is convenient to compute the net moment about a point other than the center of mass. Alternatively, by simply differentiating Eqn (9.27) we get _ HPrel ¼ _ HG þ vG=P mvG=P zfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflffl{ ¼0 þ rG=P maG=P Solving for _ HG, invoking Eqn (9.30), and using the fact that aP/G ¼ aG/P leads to MGnet ¼ _ HPrel þ rG=P maP=G (9.33) Finally, if the body is rigid, the magnitude of the position vector r of any point relative to the center of mass does not change with time. Therefore, Eqn (1.52) requires that _ r ¼ u r, leading us to conclude from Eqn (9.22) that HG ¼ Z m r ðu rÞdm Rigid body (9.34) Again, the absolute angular momentum about the center of mass depends only on the absolute angular velocity and not on the absolute translational velocity of any point of the body. 474 CHAPTER 9 Rigid Body Dynamics
  • 17. No such simplification of Eqn (9.17) exists for an arbitrary reference point P. However, if the point P is fixed in inertial space and the rigid body is rotating about P, then the position vector r from P to any point of the body is constant. It follows from Eqn (1.52) that _ r ¼ u r. According to Figure 9.8, R ¼ RP þ r Differentiating with respect to time gives _ R ¼ _ RP þ _ r ¼ 0 þ u r ¼ u r Substituting this into Eqn (9.17) yields the formula for angular momentum in this special case as HP ¼ Z m r ðu rÞdm ðRigid body rotating about fixed point PÞ (9.35) Although Eqns (9.34) and (9.35) are mathematically identical, one must keep in mind the notation of Figure 9.8. Eqn (9.35) applies only if the rigid body is in pure rotation about a stationary point in inertial space, whereas Eqn (9.34) applies unconditionally to any situation. 9.5 Moments of inertia To use Eqn (9.29) or Eqn (9.30) to solve problems, the vectors within them have to be resolved into components. To find the components of angular momentum, we must appeal to its definition. We will focus on the formula for angular momentum of a rigid body about its center of mass (Eqn (9.34)) because the expression for fixed-point rotation (Eqn (9.35)) is mathematically the same. The integrand of Eqn (9.34) can be rewritten using the bac–cab vector identity presented in Eqn (2.33), r ðu rÞ ¼ ur2 rðu$rÞ (9.36) Let the origin of a comoving xyz coordinate system be attached to the center of mass G, as shown in Figure 9.9. The unit vectors of this frame are ^ i, ^ j, and ^ k. The vectors r and u can be resolved into components in the xyz directions to get r ¼ x^ i þ y^ j þ z^ k and u ¼ ux ^ i þ uy ^ j þ uz ^ k. Substituting these vector expressions into the right side of Eqn (9.36) yields r ðu rÞ ¼ ux ^ i þ uy ^ j þ uz ^ k x2 þ y2 þ z2 x^ i þ y^ j þ z^ k uxx þ uyy þ uzz Expanding the right side and collecting the terms having the unit vectors ^ i, ^ j, and ^ k in common, we get r ðu rÞ ¼ y2 þ z2 ux xyuy xzuz ^ i þ yxux þ x2 þ z2 uy yzuz ^ j þ zxux zyuy þ x2 þ y2 uz ^ k (9.37) 9.5 Moments of inertia 475
  • 18. We put this result into the integrand of Eqn (9.34) to obtain HG ¼ Hx ^ i þ Hy ^ j þ Hz ^ k (9.38) where 8 : Hx Hy Hz 9 = ; ¼ 2 6 6 6 4 Ix Ixy Ixz Iyx Iy Iyz Izx Izy Iz 3 7 7 7 5 8 : ux uy uz 9 = ; (9.39a) or, in matrix notation, fHg ¼ ½Ifug (9.39b) The nine components of the moment of inertia matrix [I] about the center of mass are Ix ¼ R y2 þ z2 dm Ixy ¼ R xydm Ixz ¼ R xzdm Iyx ¼ R yxdm Iy ¼ R x2 þ z2 dm Iyz ¼ R yzdm Izx ¼ R zxdm Izy ¼ R zydm Iz ¼ R x2 þ y2 dm (9.40) Since Iyx ¼ Ixy, Izx ¼ Ixz, and Izy ¼ Iyz, it follows that [I] is a symmetric matrix ð½IT ¼ ½IÞ. Therefore, [I] has six independent components instead of nine. Observe that, whereas the products of inertia Ixy, Ixz, and Iyz can be positive, negative, or zero, the moments of inertia Ix, Iy, and Iz are always positive (never zero or negative) for bodies of finite dimensions. For this reason, [I] is a symmetric positive-definite matrix. Keep in mind that Eqns (9.38) and (9.39) are valid as well for axes attached to a fixed point P about which the body is rotating. The moments of inertia reflect how the mass of a rigid body is distributed. They manifest a body’s rotational inertia, that is, its resistance to being set into rotary motion or stopped once rotation is FIGURE 9.9 A comoving xyz frame used to compute the moments of inertia. 476 CHAPTER 9 Rigid Body Dynamics
  • 19. underway. It is not an object’s mass alone but how that mass is distributed that determines how the body will respond to applied torques. If the xy plane is a plane of symmetry, then for any x and y within the body there are identical mass elements located at þz and z. This means the products of inertia with z in the integrand vanish. Similar statements are true if xz or yz are symmetry planes. In summary, we conclude If the xy plane is a plane of symmetry of the body, then Ixz ¼ Iyz ¼ 0. If the xz plane is a plane of symmetry of the body, then Ixy ¼ Iyz ¼ 0. If the yz plane is a plane of symmetry of the body, then Ixy ¼ Ixz ¼ 0. It follows that if the body has two planes of symmetry relative to the xyz frame of reference, then all three products of inertia vanish, and [I] becomes a diagonal matrix such that, ½I ¼ 2 6 6 4 A 0 0 0 B 0 0 0 C 3 7 7 5 (9.41) A, B, and C are the principal moments of inertia (all positive), and the xyz axes are the body’s principal axes of inertia. In this case, relative to either the center of mass or a fixed point of rotation, we have Hx ¼ Aux Hy ¼ Buy Hz ¼ Cuz (9.42) In general, the angular velocity u and the angular momentum H are not parallel. However, if, for example, u ¼ u^ i, then according to Eqn (9.42), H ¼ Au. In other words, if the angular velocity is aligned with a principal direction, so is the angular momentum. In that case, the two vectors u and H are indeed parallel. Each of the three principal moments of inertia can be expressed as follows: A ¼ mk2 x B ¼ mk2 y C ¼ mk2 z (9.43) where m is the mass of the body and kx, ky, and kz are the three radii of gyration. One may imagine the mass of a body to be concentrated around a principal axis at a distance equal to the radius of gyration. The moments of inertia for several common shapes are listed in Figure 9.10. By symmetry, their products of inertia vanish for the coordinate axes used. Formulas for other solid geometries can be found in engineering handbooks and in dynamics textbooks. For a mass concentrated at a point, the moments of inertia in Eqn (9.40) are just the mass times the integrand evaluated at the point. That is, the moment of inertia matrix [I(m) ] of a point mass m is given by h IðmÞ i ¼ 2 6 6 4 m y2 þ z2 mxy mxz mxy m x2 þ z2 myz mxz myz m x2 þ y2 3 7 7 5 (9.44) 9.5 Moments of inertia 477
  • 20. EXAMPLE 9.4 The following table lists out the mass and coordinates of seven point masses. Find the center of mass of the system and the moments of inertia about the origin. Solution The total mass of this system is m ¼ X 7 i¼1 mi ¼ 28 kg For concentrated masses, the integral in Eqn (9.9) is replaced by the mass times its position vector. Therefore, in this case, the three components of the position vector of the center of mass are xG ¼ ð1=mÞ P7 i¼1 mixi, yG ¼ ð1=mÞ P7 i¼1 miyi, and zG ¼ ð1=mÞ P7 i¼1 mizi, so that xG ¼ 0:35 m yG ¼ 0:01964 m zG ¼ 0:4411 m Point, i Mass mi (kg) xi (m) yi (m) zi (m) 1 3 0.5 0.2 0.3 2 7 0.2 0.75 0.4 3 5 1 0.8 0.9 4 6 1.2 1.3 1.25 5 2 1.3 1.4 0.8 6 4 0.3 1.35 0.75 7 1 1.5 1.7 0.85 (b) (a) (c) FIGURE 9.10 Moments of inertia for three common homogeneous solids of mass m. (a) Solid circular cylinder. (b) Circular cylindrical shell. (c) Rectangular parallelepiped. 478 CHAPTER 9 Rigid Body Dynamics
  • 21. The total moment of inertia is the sum over all the particles of Eqn (9.44) evaluated at each point. Thus, ½I ¼ 2 4 0:39 0:3 0:45 0:3 1:02 0:18 0:45 0:18 0:87 3 5 zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{ ð1Þ þ 2 4 5:0575 1:05 0:56 1:05 1:4 2:1 0:56 2:1 4:2175 3 5 zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{ ð2Þ þ 2 4 7:25 4 4:5 4 9:05 3:6 4:5 3:6 8:2 3 5 zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{ ð3Þ þ 2 4 19:515 9:36 9 9:36 18:015 9:75 9 9:75 18:78 3 5 zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{ ð4Þ þ 2 4 5:2 3:64 2:08 3:64 4:66 2:24 2:08 2:24 7:3 3 5 zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{ ð5Þ þ 2 4 9:54 1:62 0:9 1:62 2:61 4:05 0:9 4:05 7:65 3 5 zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{ ð6Þ þ 2 4 3:6125 2:55 1:275 2:55 2:9725 1:445 1:275 1:445 5:14 3 5 zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{ ð7Þ or ½I ¼ 2 6 6 4 50:56 20:42 14:94 20:42 39:73 14:90 14:94 14:90 52:16 3 7 7 5 kg$m2 EXAMPLE 9.5 Calculate the moments of inertia of a slender, homogeneous straight rod of length [ and mass m shown in Figure 9.11. One end of the rod is at the origin and the other has coordinates (a, b, c). Solution A slender rod is one whose cross-sectional dimensions are negligible compared with its length. The mass is concentrated along its centerline. Since the rod is homogeneous, the mass per unit length r is uniform and given by r ¼ m ‘ (a) The length of the rod is ‘ ¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi a2 þ b2 þ c2 p (a, b, c) (0, 0, 0) A B FIGURE 9.11 A uniform slender bar of mass m and length [. 9.5 Moments of inertia 479
  • 22. Starting with Ix, we have from Eqn (9.40), Ix ¼ Z‘ 0 y2 þ z2 rds in which we replaced the element of mass dm by rds, where ds is the element of length along the rod. The distance s is measured from end A of the rod, so that the x, y, and z coordinates of any point along it are found in terms of s by the following relations: x ¼ s ‘ a y ¼ s ‘ b z ¼ s ‘ c Thus, Ix ¼ Z‘ 0 s2 ‘2 b2 þ s2 ‘2 c2 rds ¼ r b2 þ c2 ‘2 Z‘ 0 s2 ds ¼ 1 3 r b2 þ c2 ‘ Substituting Eqn (a) yields Ix ¼ 1 3 m b2 þ c2 In precisely the same way, we find Iy ¼ 1 3 m a2 þ c2 Iz ¼ 1 3 m a2 þ b2 For Ixy we have Ixy ¼ Z‘ 0 xyrds ¼ Z‘ 0 s ‘ a $ s ‘ brds ¼ r ab ‘2 Z‘ 0 s2 ds ¼ 1 3 rab‘ Once again using Eqn (a), Ixy ¼ 1 3 mab Likewise, Ixz ¼ 1 3 mac Iyz ¼ 1 3 mbc EXAMPLE 9.6 The gyro rotor (Figure 9.12) in Example 9.3 has a mass m of 5 kg, radius r of 0.08 m, and thickness t of 0.025 m. If N ¼ 2.1 rad/s, _ q ¼ 4 rad=s, u ¼ 10.5 rad/s, and q ¼ 60, calculate (a) The angular momentum of the rotor about its center of mass G in the body-fixed xyz frame. (b) The angle between the rotor’s angular velocity vector and its angular momentum vector. 480 CHAPTER 9 Rigid Body Dynamics
  • 23. Solution Equation (f) from Example 9.3 gives the components of the absolute angular velocity of the rotor in the moving xyz frame. ux ¼ _ q ¼ 4 rad=s uy ¼ N sin q ¼ 2:1$sin 60 ¼ 1:819 rad=s uz ¼ uspin þ N cos q ¼ 10:5 þ 2:1$cos 60 ¼ 11:55 rad=s (a) Therefore, u ¼ 4^ i þ 1:819^ j þ 11:55^ k ðrad=sÞ (b) All three coordinate planes of the body-fixed xyz frame contain the center of mass G and all are planes of symmetry of the circular cylindrical rotor. Therefore, [xy ¼ [zx ¼ [yz ¼ 0. From Figure 9.10(a), we see that the nonzero diagonal entries in the moment of inertia tensor are A ¼ B ¼ 1 12 mt2 þ 1 4 mr2 ¼ 1 12 $5$0:0252 þ 1 4 $5$0:082 ¼ 0:008260 kg$m2 C ¼ 1 2 mr2 ¼ 1 2 $5$0:082 ¼ 0:0160 kg$m2 (c) We can use Eqn (9.42) to calculate the angular momentum, because the origin of the xyz frame is the rotor’s center of mass (which in this case also happens to be a fixed point of rotation, which is another reason why we can use Eqn (9.42)). Substituting Eqns (a) and (c) in Eqn (9.42) yields Hx ¼ Aux ¼ 0:008260$4 ¼ 0:03304 kg$m2=s Hy ¼ Buy ¼ 0:008260$1:819 ¼ 0:0150 kg$m2=s Hz ¼ Cuz ¼ 0:0160$11:55 ¼ 0:1848 kg$m2=s (d) so that H ¼ 0:03304^ i þ 0:0150^ j þ 0:1848^ k kg$m2=s (e) θ ω FIGURE 9.12 Rotor of the gyroscope in Figure 9.4. 9.5 Moments of inertia 481
  • 24. The angle f between H and u is found by taking the dot product of the two vectors, f ¼ cos1 H$u Hu ¼ cos1 2:294 0:1883$12:36 ¼ 9:717 (f) As this problem illustrates, the angular momentum and the angular velocity are in general not collinear. Consider a coordinate system x0y0z0 with the same origin as xyz but a different orientation. Let [Q] be the orthogonal matrix ð½Q1 ¼ ½QT Þ that transforms the components of a vector from the xyz system to the x0y0z0 frame. Recall from Section 4.5 that the rows of [Q] are the direction cosines of the x0y0z0 axes relative to xyz. If fH0 g comprises the components of the angular momentum vector along the x0y0z0 axes, then fH0 g is obtained from its components {H} in the xyz frame by the relation fH0 g ¼ ½QfHg From Eqn (9.39), we can write this as fH0 g ¼ ½Q½Ifug (9.45) where [I] is the moment of inertia matrix (Eqn (9.39)) in the xyz coordinates. Like the angular mo- mentum vector, the components {u} of the angular velocity vector in the xyz system are related to those in the primed system ðfu0gÞ by the expression fu0 g ¼ ½Qfug The inverse relation is simply fug ¼ ½Q1 fu0 g ¼ ½QT fu0 g (9.46) Substituting this into Eqn (9.45), we get fH0 g ¼ ½Q½I½QT fu0 g (9.47) But the components of angular momentum and angular velocity in the x0y0z0 frame are related by an equation of the same form as Eqn (9.39), so that fH0 g ¼ ½I0 fu0 g (9.48) where ½I0 comprises the components of the inertia matrix in the primed system. Comparing the right- hand sides of Eqns (9.47) and (9.48), we conclude that ½I0 ¼ ½Q½I½QT (9.49a) that is, 2 6 6 4 Ix0 Ix0y0 Ix0z0 Iy0x0 Iy0 Iy0z0 Iz0x0 Iz0y0 Iz0 3 7 7 5 ¼ 2 6 6 4 Q11 Q12 Q13 Q21 Q22 Q23 Q31 Q32 Q33 3 7 7 5 2 6 6 4 Ix Ixy Ixz Iyx Iy Iyz Izx Izy Iz 3 7 7 5 2 6 6 4 Q11 Q21 Q31 Q12 Q22 Q32 Q13 Q32 Q33 3 7 7 5 (9.49b) 482 CHAPTER 9 Rigid Body Dynamics
  • 25. This shows how to transform the components of the inertia matrix from the xyz coordinate system to any other orthogonal system with a common origin. Thus, for example, Ix0 ¼ P Q11 Q12 Q13 R zfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflffl{ PRow 1R 2 6 6 4 Ix Ixy Ixz Iyx Iy Iyz Izx Izy Iz 3 7 7 5 8 : Q11 Q12 Q13 9 = ; zfflfflfflfflffl}|fflfflfflfflffl{ PRow 1R T Iy0z0 ¼ P Q21 Q22 Q23 R zfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflffl{ PRow 2R 2 6 6 4 Ix Ixy Ixz Iyx Iy Iyz Izx Izy Izz 3 7 5 8 : Q31 Q32 Q33 9 = ; zfflfflfflfflffl}|fflfflfflfflffl{ PRow 3R T (9.50) etc. Any object represented by a square matrix whose components transform according to Eqn (9.49) is called a second-order tensor. We may therefore refer to [I] as the inertia tensor. EXAMPLE 9.7 Find the mass moment of inertia of the system of point masses in Example 9.4 about an axis from the origin through the point with coordinates (2 m, 3 m, 4 m). Solution From Example 9.4, the moment of inertia tensor for the system of point masses is ½I ¼ 2 6 6 4 50:56 20:42 14:94 20:42 39:73 14:90 14:94 14:90 52:16 3 7 7 5 kg$m2 The vector V connecting the origin with (2 m, 3 m, 4 m) is V ¼ 2^ i 3^ j þ 4^ k The unit vector in the direction of V is ^ uV ¼ V kVk ¼ 0:3714^ i 0:5571^ j þ 0:7428^ k We may consider ^ uV as the unit vector along the x0-axis of a rotated Cartesian coordinate system. Then, from Eqn (9.50) we get ‘V 0 ¼ P 0:3714 0:5571 0:7428 R 2 6 6 6 6 4 50:56 20:42 14:94 20:42 39:73 14:90 14:94 14:90 52:16 3 7 7 7 7 5 8 : 0:3714 0:5571 0:7428 9 = ; ¼ P 0:3714 0:5571 0:7428 R 8 : 3:695 3:482 24:90 9 = ; ¼ 19:06 kg$m2 9.5 Moments of inertia 483
  • 26. EXAMPLE 9.8 For the satellite of Example 9.2, which is reproduced in Figure 9.13, the data are as follows: N ¼ 0.1 rad/s and _ q ¼ 0:01 rad=s, in the directions shown: q ¼ 40 and d0 ¼ 1.5 m. The length, width, and thickness of the panel are [ ¼ 6 m, w ¼ 2 m, and t ¼ 0.025 m. The uniformly distributed mass of the panel is 50 kg. Find the angular momentum of the panel relative to the center of mass O of the satellite. Solution We can treat the panel as a thin parallelepiped. The panel’s xyz axes have their origin at the center of mass G of the panel and are parallel to its three edge directions. According to Figure 9.10(c), the moments of inertia relative to the xyz coordinate system are ‘Gx ¼ 1 12 m ‘2 þ t2 ¼ 1 12 $50$ 62 þ 0:0252 ¼ 150:0 kg$m2 ‘Gy ¼ 1 12 m w2 þ t2 ¼ 1 12 $50$ 22 þ 0:0252 ¼ 16:67 kg$m2 ‘Gz ¼ 1 12 m w2 þ ‘2 ¼ 1 12 $50$ 22 þ 62 ¼ 166:7 kg$m2 ‘Gxy ¼ ‘Gxz ¼ ‘Gyz ¼ 0 (a) In matrix notation, ½lG ¼ 2 6 6 4 150:0 0 0 0 16:67 0 0 0 166:7 3 7 7 5 kg$m2 (b) The unit vectors of the satellite’s x0y0z0 system are related to those of the panel’s xyz frame. By inspection, ^ i 0 ¼ sin q^ i þ cos q^ k ¼ 0:6428^ i þ 0:7660^ k ^ j 0 ¼ ^ j ^ k 0 ¼ cos q^ i þ sin q^ k ¼ 0:7660^ i þ 0:6428^ k (c) θ FIGURE 9.13 A satellite and solar panel. 484 CHAPTER 9 Rigid Body Dynamics
  • 27. The matrix [Q] of the transformation from xyz to x0y0z0 comprises the direction cosines of ^ i 0 , ^ j 0 , and ^ k 0 : ½Q ¼ 2 6 6 4 0:6428 0 0:7660 0 1 0 0:7660 0 0:6428 3 7 7 5 (d) In Example 9.2, we found that the absolute angular velocity of the panel, in the satellite’s x0y0z0 frame of reference, is u ¼ _ q^ j 0 þ N^ k 0 ¼ 0:01^ j 0 þ 0:1^ k 0 ðrad=sÞ That is, fu0 g ¼ 8 : 0 0:01 0:1 9 = ; ðrad=sÞ (e) To find the absolute angular momentum fH0 Gg in the satellite system requires the use of Eqn (9.39), fH0 Gg ¼ ½I0 Gfu0 g (f) Before doing so, we must transform the components of the moments of the inertia tensor in Eqn (b) from the unprimed system to the primed system, by means of Eqn (9.49), ½I0 G ¼ ½Q½IG½QT ¼ 2 6 6 4 0:6428 0 0:7660 0 1 0 0:7660 0 0:6428 3 7 7 5 2 6 6 4 150:0 0 0 0 16:67 0 0 0 166:7 3 7 7 5 2 6 6 4 0:6428 0 0:7660 0 1 0 0:7660 0 0:6428 3 7 7 5 so that ½I0 G ¼ 2 6 6 4 159:8 0 8:205 0 16:67 0 8:205 0 156:9 3 7 7 5 kg$m2 (g) Then Eqn (f) yields fH0 Gg ¼ 2 6 6 4 159:8 0 8:205 0 16:67 0 8:205 0 156:9 3 7 7 5 8 : 0 0:01 0:1 9 = ; ¼ 8 : 0:8205 0:1667 15:69 9 = ; kg$m2 =s or, in vector notation, HG ¼ 0:8205^ i 0 0:1667^ j 0 þ 15:69^ k 0 kg$m2 =s (h) This is the absolute angular momentum of the panel about its own center of mass G, and it is used in Eqn (9.27) to calculate the angular momentum HOrel relative to the satellite’s center of mass O, HOrel ¼ HG þ rG=O mvG=O (i) rG/O is the position vector from O to G, rG=O ¼ dO þ ‘ 2 ^ j 0 ¼ 1:5 þ 6 2 ^ j 0 ¼ 4:5^ j 0 ðmÞ (j) 9.5 Moments of inertia 485
  • 28. The velocity of G relative to O, vG/O, is found from Eqn (9.2), vG=O ¼ usatellite rG=O ¼ N^ k 0 rG=O ¼ 0:1^ k 0 4:5^ j 0 ¼ 0:45^ i 0 ðm=sÞ (k) Substituting Eqns (h), (j) and (k) into Eqn (i) finally yields HOrel ¼ 0:8205^ i 0 0:1667^ j 0 þ 15:69^ k 0 þ 4:5^ j 0 h 50 0:45^ i 0 i ¼ 0:8205^ i 0 0:1667^ j 0 þ 116:9^ k 0 kg$m2=s (l) Note that we were unable to use Eqn (9.21) to find the absolute angular momentum HO because that requires knowing the absolute velocity vG, which in turn depends on the absolute velocity of O, which was not provided. How can we find the direction cosine matrix [Q] such that Eqn (9.49) will yield a moment of inertia matrix ½I0 that is diagonal, that is, of the form given by Eqn (9.41)? In other words, how do we find the principal directions (“eigenvectors”) and the corresponding principal values (“eigenvalues”) of the moment of inertia tensor? Let the angular velocity vector u be parallel to the principal direction defined by the vector e, so that u ¼ be, where b is a scalar. Since u points in the principal direction of the inertia tensor, so must H, which means H is also parallel to e. Therefore, H ¼ ae, where a is a scalar. From Eqn (9.39), it follows that afeg ¼ ½IðbfegÞ or ½Ifeg ¼ lfeg where l ¼ a/b (a scalar). That is, 2 4 Ix Ixy Ixz Ixy Iy Iyz Ixz Iyz Iz 3 5 8 : ex ey ez 9 = ; ¼ l 8 : ex ey ez 9 = ; This can be written 2 4 Ix l Ixy Ixz Ixy Iy l Iyz Ixz Iyz Iz l 3 5 8 : ex ey ez 9 = ; ¼ 8 : 0 0 0 9 = ; (9.51) The trivial solution of Eqn (9.51) is e ¼ 0, which is of no interest. The only way that Eqn (9.51) will not yield the trivial solution is if the coefficient matrix on the left is singular. That will occur if its determinant vanishes, that is, if Ix l Ixy Ixz Ixy Iy l Iyz Ixz Iyz Iz l ¼ 0 (9.52) 486 CHAPTER 9 Rigid Body Dynamics
  • 29. Expanding the determinant, we find Ix l Ixy Ixz Ixy Iy l Iyz Ixz Iyz Iz l ¼ l3 þ J1l2 J2l þ J3 (9.53) where J1 ¼ Ix þ Iy þ Iz J2 ¼ Ix Ixy Ixy Iy þ Ix Ixz Ixz Iz þ Iy Iyz Iyz Iz J3 ¼ Ix Ixy Ixz Ixy Iy Iyz Ixz Iyz Iz (9.54) J1, J2, and J3 are invariants, that is, they have the same value in every Cartesian coordinate system. Equations (9.52) and (9.53) yield the characteristic equation of the tensor [I], l3 J1l2 þ J2l J3 ¼ 0 (9.55) The three roots lp (p ¼ 1, 2, 3) of this cubic equation are real, since [I] is symmetric; furthermore, they are all positive, since [I] is a positive-definite matrix. We substitute each root, or eigenvalue, lp back into Eqn (9.51) to obtain 2 6 6 4 Ix lp Ixy Ixz Ixy Iy lp Iyz Ixz Iyz Iz lp 3 7 7 5 8 : e ðpÞ x e ðpÞ y e ðpÞ z 9 = ; ¼ 8 : 0 0 0 9 = ; ; p ¼ 1; 2; 3 (9.56) Solving this system yields the three eigenvectors e(p) corresponding to each of the three eigenvalues lp. The three eigenvectors are orthogonal, also due to the symmetry of the matrix [I]. Each eigenvalue is a principal moment of inertia, and its corresponding eigenvector is a principal direction. EXAMPLE 9.9 Find the principal moments of inertia and the principal axes of inertia of the inertia tensor. ½I ¼ 2 6 6 4 100 20 100 20 300 50 100 50 500 3 7 7 5 kg$m2 9.5 Moments of inertia 487
  • 30. Solution We seek the nontrivial solutions of the system [I]{e} ¼ l{e}, that is, 2 6 6 4 100 l 20 100 20 300 l 50 100 50 500 l 3 7 7 5 8 : ex ey ez 9 = ; ¼ 8 : 0 0 0 9 = ; (a) From Eqn (9.54), J1 ¼ 100 þ 300 þ 500 ¼ 900 J2 ¼ 100 20 20 300 þ 100 100 100 500 þ 300 50 50 500 ¼ 217; 100 J3 ¼ 100 20 100 20 300 50 100 50 500 ¼ 11; 350; 000 (b) Thus, the characteristic equation is l3 900l2 þ 217;100l 11;350;000 ¼ 0 (c) The three roots are the principal moments of inertia, which are found to be l1 ¼ 532:052 l2 ¼ 295:840 l3 ¼ 72:1083 (d) Each of these is substituted, in turn, back into Eqn (a) to find its corresponding principal direction. Substituting l1 ¼ 532:052 kg$m2 into Eqn (a) we obtain 2 6 6 4 432:052 20:0000 100:0000 20:0000 232:052 50:0000 100:0000 50:0000 32:0519 3 7 7 5 8 : e ð1Þ x e ð1Þ y e ð1Þ z 9 = ; ¼ 8 : 0 0 0 9 = ; (e) Since the determinant of the coefficient matrix is zero, at most two of the three equations in Eqn (e) are inde- pendent. Thus, at most, two of the three components of the vector e(1) can be found in terms of the third. We can therefore arbitrarily set e ð1Þ x ¼ 1 and solve for e ð1Þ y and e ð1Þ z using any two of the independent equations in Eqn (e). With e ð1Þ x ¼ 1, the first two of Eqn (e) become 20:0000e ð1Þ y 100:000e ð1Þ z ¼ 432:052 232:052e ð1Þ y 50:000e ð1Þ z ¼ 20:0000 (f) Solving these two equations for e ð1Þ y and e ð1Þ z yields, together with the assumption that e ð1Þ x ¼ 1, e ð1Þ x ¼ 1:00000 e ð1Þ y ¼ 0:882793 e ð1Þ z ¼ 4:49708 (g) The unit vector in the direction of e(1) is ^ e1 ¼ eð1Þ eð1Þ ¼ 1:00000^ i þ 0:882793^ j 4:49708^ k ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1:000002 þ 0:8827932 þ ð4:49708Þ2 q 488 CHAPTER 9 Rigid Body Dynamics
  • 31. or ^ e1 ¼ 0:213186^ i þ 0:188199^ j 0:958714^ k l1 ¼ 532:052 kg$m2 (h) Substituting l2 ¼ 295:840 kg$m2 into Eqn (a) and proceeding as above we find that ^ e2 ¼ 0:176732^ i 0:972512^ j 0:151609^ k l2 ¼ 295:840 kg$m2 (i) The two unit vectors ^ e1 and ^ e2 define two of the three principal directions of the inertia tensor. Observe that ^ e1$^ e2 ¼ 0, as must be the case for symmetric matrices. To obtain the third principal direction ^ e3, we can substitute l3 ¼ 72:1083 kg$m2 into Eqn (a) and pro- ceed as above. However, since the inertia tensor is symmetric, we know that the three principal directions are mutually orthogonal, which means ^ e3 ¼ ^ e1 ^ e2. Substituting Eqns (h) and (i) into the crossproduct, we find that ^ e3 ¼ 0:960894^ i 0:137114^ j 0:240587^ k l3 ¼ 72:1083 kg$m2 (j) We can check our work by substituting l3 and ^ e3 into Eqn (a) and verify that it is indeed satisfied: 2 6 6 4 100 72:1083 20 100 20 300 72:1083 50 100 50 500 72:1083 3 7 7 5 8 : 0:960894 0:137114 0:240587 9 = ; ¼ O 8 : 0 0 0 9 = ; (k) The components of the vectors ^ e1, ^ e2, and ^ e3 define the three rows of the orthogonal transformation [Q] from the xyz system into the x0y0z0 system aligned along the three principal directions: ½Q ¼ 2 6 6 4 0:213186 0:188199 0:958714 0:176732 0:972512 0:151609 0:960894 0:137114 0:240587 3 7 7 5 (l) Indeed, if we apply the transformation in Eqn (9.49), ½I0 ¼ ½Q½I½QT , we find ½I0 ¼ 2 6 6 4 0:213186 0:188199 0:958714 0:176732 0:972512 0:151609 0:960894 0:137114 0:240587 3 7 7 5 2 6 6 4 100 20 100 20 300 50 100 50 500 3 7 7 5 2 6 6 4 0:213186 0:176732 0:960894 0:188199 0:972512 0:137114 0:958714 0:151609 0:240587 3 7 7 5 ¼ 2 6 6 4 532:052 0 0 0 295:840 0 0 0 72:1083 3 7 7 5 kg$m2 9.5 Moments of inertia 489
  • 32. An alternative to the above hand calculations in Example 9.9 is to type the following lines in the MATLAB Command Window: I ¼ [ 100 -20 -100 -20 300 -50 -100 -50 500]; [eigenVectors, eigenValues] ¼ eig(I) Hitting the Enter (or Return) key yields the following output to the Command Window: eigenVectors ¼ 0.9609 0.1767 -0.2132 0.1371 -0.9725 -0.1882 0.2406 -0.1516 0.9587 eigenValues ¼ 72.1083 0 0 0 295.8398 0 0 0 532.0519 Two of the eigenvectors delivered by MATLAB are opposite in direction to those calculated in Example 9.9. This illustrates the fact that we can determine an eigenvector only to within an arbitrary scalar factor. That is, suppose e is an eigenvector of the tensor [I] so that [I]{e} ¼ l{e}. Multiplying this equation through by an arbitrary scalar a yields [I]{e}a ¼ l{e}a, or [I]{ae} ¼ l{ae}, which means that {ae} is an eigenvector corresponding to the same eigenvalue l. Parallel axis theorem Suppose the rigid body in Figure 9.14 is in pure rotation about point P. Then, according to Eqn (9.39), fHPrel g ¼ ½IPfug (9.57) η ζ ξ FIGURE 9.14 The moments of inertia are to be computed at P, given their values at G. 490 CHAPTER 9 Rigid Body Dynamics
  • 33. where [IP] is the moment of inertia tensor about P, given by Eqn (9.40) with x ¼ xG=P þ x y ¼ yG=P þ h z ¼ zG=P þ z On the other hand, we have from Eqn (9.27) that HPrel ¼ HG þ rG=P mvG=P (9.58) The vector rG/P mvG/P is the angular momentum about P of the concentrated mass m located at the center of mass G. Using matrix notation, it is computed as follows: n rG=P mvG=P o hfH ðmÞ Prel g ¼ h I ðmÞ P i fug (9.59) where h I ðmÞ P i , the moment of inertia of the point mass m about P, is obtained from Eqn (9.44), with x ¼ xG/P, y ¼ yG/P, and z ¼ zG/P. That is, ½I ðmÞ P ¼ 2 6 6 6 4 m y2 G=P þ z2 G=P mxG=P yG=P mxG=P zG=P mxG=P yG=P m x2 G=P þ z2 G=P myG=P zG=P mxG=P zG=P myG=P zG=P m x2 G=P þ y2 G=P 3 7 7 7 5 (9.60) Of course, Eqn (9.39) requires fHGg ¼ ½IGfug Substituting this together with Eqns (9.57) and (9.59) into Eqn (9.58) yields ½IPfug ¼ ½IGfug þ h I ðmÞ P i fug ¼ ½IG þ h I ðmÞ P i fug From this, we may infer the parallel axis theorem, ½IP ¼ ½IG þ h I ðmÞ P i (9.61) The moment of inertia about P is the moment of inertia about the parallel axes through the center of mass plus the moment of inertia of the center of mass about P. That is, IPx ¼ IGx þ m y2 G=P þ z2 G=P IPy ¼ IGy þ m y2 G=P þ x2 G=P IPz ¼ IGz þ m x2 G=P þ y2 G=P IPxy ¼ IGxy mxG=P yG=P IPxz ¼ IGxz mxG=P zG=P IPyz ¼ IGyz myG=P zG=P (9.62) 9.5 Moments of inertia 491
  • 34. EXAMPLE 9.10 Find the moments of inertia of the rod in Example 9.5 (Figure 9.15) about its center of mass G. Solution From Example 9.5, ½IA ¼ 2 6 6 6 6 6 6 6 6 6 4 1 3 m b2 þ c2 1 3 mab 1 3 mac 1 3 mab 1 3 m a2 þ c2 1 3 mbc 1 3 mac 1 3 mbc 1 3 m a2 þ b2 3 7 7 7 7 7 7 7 7 7 5 Using Eqn (9.62)1, and noting the coordinates of the center of mass in Figure 9.15, IGx ¼ IAx m h ðyG 0Þ2 þ ðzG 0Þ2 i ¼ 1 3 m b2 þ c2 m b 2 2 þ c 2 2 ¼ 1 12 m b2 þ c2 Equation (9.62)4 yields IGxy ¼ IAxy þ mðxG 0ÞðyG 0Þ ¼ 1 3 mab þ m$ a 2 $ b 2 ¼ 1 12 mab The remaining four moments of inertia are found in a similar fashion, so that ½IG ¼ 2 6 6 6 6 6 6 6 6 6 6 4 1 12 m b2 þ c2 1 12 mab 1 12 mac 1 12 mab 1 12 m a2 þ c2 1 12 mbc 1 12 mac 1 12 mbc 1 12 m a2 þ b2 3 7 7 7 7 7 7 7 7 7 7 5 (9.63) FIGURE 9.15 A uniform slender rod. 492 CHAPTER 9 Rigid Body Dynamics
  • 35. EXAMPLE 9.11 Calculate the principal moments of inertia about the center of mass and the corresponding principal directions for the bent rod in Figure 9.16. Its mass is uniformly distributed at 2 kg/m. Solution The mass of each of the rod segments is m1 ¼ 2$0:4 ¼ 0:8 kg m2 ¼ 2$0:5 ¼ 1 kg m3 ¼ 2$0:3 ¼ 0:6 kg m4 ¼ 2$0:2 ¼ 0:4 kg (a) The total mass of the system is m ¼ X 4 i¼1 mi ¼ 2:8 kg (b) The coordinates of each segment’s center of mass are xG1 ¼ 0 yG1 ¼ 0 zG1 ¼ 0:2 m xG2 ¼ 0 yG2 ¼ 0:25 m zG2 ¼ 0:2 m xG3 ¼ 0:15 m yG3 ¼ 0:5 m zG3 ¼ 0 xG4 ¼ 0:3 m yG4 ¼ 0:4 m zG4 ¼ 0 (c) If the slender rod in Figure 9.15 is aligned with, say, the x-axis, then a ¼ [ and b ¼ c ¼ 0, so that according to Eqn (9.63), ½IG ¼ 2 6 6 6 6 6 4 0 0 0 0 1 12 m‘2 0 0 0 1 12 m‘2 3 7 7 7 7 7 5 That is, the moment of inertia of a slender rod about the axes normal to the rod at its center of mass is 1 12 m‘2, where m and [ are the mass and length of the rod, respectively. Since the mass of a slender bar is assumed to be concentrated along the axis of the bar (its cross-sectional dimensions are infinitesimal), the moment of inertia 4 3 2 1 0.4 m 0.5 m 0.3 m 0.2 m O FIGURE 9.16 Bent rod for which the principal moments of inertia are to be determined. 9.5 Moments of inertia 493
  • 36. about the centerline is zero. By symmetry, the products of inertia about the axes through the center of mass are all zero. Using this information and the parallel axis theorem, we find the moments and products of inertia of each rod segment about the origin O of the xyz system as follows: Rod 1: I ð1Þ x ¼ I ð1Þ G1 x þ m1 y2 G1 þ z2 G1 ¼ 1 12 $0:8$0:42 þ h 0:8 0 þ 0:22 i ¼ 0:04267 kg-m2 I ð1Þ y ¼ I ð1Þ G1 y þ m1 x2 G1 þ z2 G1 ¼ 1 12 $0:8$0:42 þ h 0:8 0 þ 0:22 i ¼ 0:04267 kg-m2 I ð1Þ z ¼ I ð1Þ G1 z þ m1 x2 G1 þ y2 G1 ¼ 0 þ ½0:8ð0 þ 0Þ ¼ 0 I ð1Þ xy ¼ I ð1Þ G1 xy m1xG1 yG1 ¼ 0 ½0:8ð0Þð0Þ ¼ 0 I ð1Þ xz ¼ I ð1Þ G1 xz m1xG1 zG1 ¼ 0 ½0:8ð0Þð0:2Þ ¼ 0 I ð1Þ yz ¼ I ð1Þ G1 yz m1yG1 zG1 ¼ 0 ½0:8ð0Þð0Þ ¼ 0 Rod 2: I ð2Þ x ¼ I ð2Þ G2 x þ m2 y2 G2 þ z2 G2 ¼ 1 12 $1:0$0:52 þ h 1:0 0 þ 0:252 i ¼ 0:08333 kg-m2 I ð2Þ y ¼ I ð2Þ G2 y þ m2 x2 G2 þ z2 G2 ¼ 0 þ ½1:0ð0 þ 0Þ ¼ 0 I ð2Þ z ¼ I ð2Þ G2 z þ m2 x2 G2 þ y2 G2 ¼ 1 12 $1:0$0:52 þ h 1:0 0 þ 0:52 i ¼ 0:08333 kg-m2 I ð2Þ xy ¼ I ð2Þ G2 xy m2xG2 yG2 ¼ 0 ½1:0ð0Þð0:5Þ ¼ 0 I ð2Þ xz ¼ I ð2Þ G2 xz m2xG2 zG2 ¼ 0 ½1:0ð0Þð0Þ ¼ 0 I ð2Þ yz ¼ I ð2Þ G2 yz m2yG2 zG2 ¼ 0 ½1:0ð0:5Þð0Þ ¼ 0 Rod 3: I ð3Þ x ¼ I ð3Þ G3 x þ m3 y2 G3 þ z2 G3 ¼ 0 þ 0:6 0:52 þ 0 ¼ 0:15 kg-m2 I ð3Þ y ¼ I ð3Þ G3 y þ m3 x2 G3 þ z2 G3 ¼ 1 12 $0:6$0:32 þ h 0:6 0:152 þ 0 i ¼ 0:018 kg-m2 I ð3Þ z ¼ I ð3Þ G3 z þ m3 x2 G3 þ y2 G3 ¼ 1 12 $0:6$0:32 þ h 0:6 0:152 þ 0:52 i ¼ 0:1680 kg-m2 I ð3Þ xy ¼ I ð3Þ G3 xy m3xG3 yG3 ¼ 0 ½0:6ð0:15Þð0:5Þ ¼ 0:045 kg-m2 I ð3Þ xz ¼ I ð3Þ G3 xz m3xG3 zG3 ¼ 0 ½0:6ð0:15Þð0Þ ¼ 0 I ð3Þ yz ¼ I ð3Þ G3 yz m3yG3 zG3 ¼ 0 ½0:6ð0:5Þð0Þ ¼ 0 494 CHAPTER 9 Rigid Body Dynamics
  • 37. Rod 4: I ð4Þ x ¼ I ð4Þ G4 x þ m4 y2 G4 þ z2 G4 ¼ 1 12 $0:4$0:22 þ h 0:4 0:42 þ 0 i ¼ 0:06533 kg-m2 I ð4Þ y ¼ I ð4Þ G4 y þ m4 x2 G4 þ z2 G4 ¼ 0 þ 0:4 0:32 þ 0 ¼ 0:0360 kg-m2 I ð4Þ z ¼ I ð4Þ G4 z þ m4 x2 G4 þ y2 G4 ¼ 1 12 $0:4$0:22 þ h 0:4 0:32 þ 0:42 i ¼ 0:1013 kg-m2 I ð4Þ xy ¼ I ð4Þ G4 xy m4xG4 yG4 ¼ 0 ½0:4ð0:3Þð0:4Þ ¼ 0:0480 kg-m2 I ð4Þ xz ¼ I ð4Þ G4 xz m4xG4 zG4 ¼ 0 ½0:4ð0:3Þð0Þ ¼ 0 I ð4Þ yz ¼ I ð4Þ G4 yz m4yG4 zG4 ¼ 0 ½0:4ð0:4Þð0Þ ¼ 0 The total moments of inertia for all the four rods about O are Ix ¼ P 4 i¼1 I ðiÞ x ¼ 0:3413 kg-m2 Iy ¼ P 4 i¼1 I ðiÞ y ¼ 0:09667 kg-m2 Iz ¼ P 4 i¼1 I ðiÞ z ¼ 0:3527 kg-m2 Ixy ¼ P 4 i¼1 I ðiÞ xy ¼ 0:0930 kg-m2 Ixz ¼ P 4 i¼1 I ðiÞ xz ¼ 0 Iyz ¼ P 4 i¼1 I ðiÞ yz ¼ 0 (d) The coordinates of the center of mass of the system of four rods are, from Eqns (a)e(c), xG ¼ 1 m X 4 i¼1 mixGi ¼ 1 2:8 $0:21 ¼ 0:075 m yG ¼ 1 m X 4 i¼1 miyGi ¼ 1 2:8 $0:71 ¼ 0:2536 m zG ¼ 1 m X 4 i¼1 mizGi ¼ 1 2:8 $0:16 ¼ 0:05714 m (e) We use the parallel axis theorems to shift the moments of inertia in Eqn (d) to the center of mass G of the system. IGx ¼ Ix m y2 G þ z2 G ¼ 0:3413 0:1892 ¼ 0:1522 kg-m2 IGy ¼ Iy m x2 G þ z2 G ¼ 0:09667 0:02489 ¼ 0:07177 kg-m2 IGz ¼ Iz m x2 G þ y2 G ¼ 0:3527 0:1958 ¼ 0:1569 kg-m2 IGxy ¼ Ixy þ mxGyG ¼ 0:093 þ 0:05325 ¼ 0:03975 kg-m2 IGxz ¼ Ixz þ mxGzG ¼ 0 þ 0:012 ¼ 0:012 kg-m2 IGyz ¼ Iyz þ myGzG ¼ 0 þ 0:04057 ¼ 0:04057 kg-m2 9.5 Moments of inertia 495
  • 38. Therefore, the inertia tensor, relative to the center of mass, is ½I ¼ 2 6 6 4 IGx IGxy IGxz IGxy IGy IGyz IGxz IGyz IGz 3 7 7 5 ¼ 2 6 6 4 0:1522 0:03975 0:012 0:03975 0:07177 0:04057 0:012 0:04057 0:1569 3 7 7 5 kg$m2 (f) To find the three principal moments of inertia, we may proceed as in Example 9.9, or simply enter the following lines in the MATLAB Command Window: I ¼ [ 0.1522 -0.03975 0.012 -0.03975 0.07177 0.04057 0.012 0.04057 0.1569 ]; [eigenVectors, eigenValues] ¼ eig(IG) to obtain eigenVectors ¼ 0.3469 -0.8482 -0.4003 0.8742 0.1378 0.4656 -0.3397 -0.5115 0.7893 eigenValues ¼ 0.0402 0 0 0 0.1658 0 0 0 0.1747 Hence, the three principal moments of inertia and their principal directions are l1 ¼ 0:04023 kg-m2 eð1Þ ¼ 0:3469^ i þ 0:8742^ j 0:3397^ k l2 ¼ 0:1658 kg-m2 eð2Þ ¼ 0:8482^ i þ 0:1378^ j 0:5115^ k l3 ¼ 0:1747 kg-m2 eð3Þ ¼ 0:4003^ i þ 0:4656^ j þ 0:7893^ k 9.6 Euler’s equations For either the center of mass G or for a fixed point P about which the body is in pure rotation, we know from Eqns (9.29) and (9.30) that Mnet ¼ _ H (9.64) Using a comoving coordinate system, with angular velocity U and its origin located at the point (G or P), the angular momentum has the analytical expression H ¼ Hx ^ i þ Hy ^ j þ Hz ^ k (9.65) We shall henceforth assume, for simplicity, that ðaÞ The moving xyz axes are the principal axes of inertia and (9.66a) ðbÞ The moments of inertia relative to xyz are constant in time: (9.66b) 496 CHAPTER 9 Rigid Body Dynamics
  • 39. Equations (9.42) and (9.66a) imply that H ¼ Aux ^ i þ Buy ^ j þ Cuz ^ k (9.67) where A, B, and C are the principal moments of inertia. According to Eqn (1.56), the time derivative of H is _ H ¼ _ HÞrel þ U H, so that Eqn (9.64) can be written as Mnet ¼ _ H rel þ U H (9.68) Keep in mind that, whereas U (the angular velocity of the moving xyz coordinate system) and u (the angular velocity of the rigid body itself) are both absolute kinematic quantities, Eqn (9.68) contains their components as projected onto the axes of the noninertial xyz frame given by u ¼ ux ^ i þ uy ^ j þ uz ^ k U ¼ Ux ^ i þ Uy ^ j þ Uz ^ k The absolute angular acceleration a is obtained using Eqn (1.56) as a ¼ _ u ¼ dux dt ^ i þ duy dt ^ j þ duz dt ^ k zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{ arel þ U u that is, a ¼ _ ux þ Uyuz Uzuy ^ i þ _ uy þ Uzux Uxuz ^ j þ _ uz þ Uxuy Uyux ^ k (9.69) Clearly, it is generally true that axs _ ux ays _ uy azs _ uz From Eqn (1.57) and Eqn (9.67), _ H rel ¼ dðAuxÞ dt ^ i þ d Buy dt ^ j þ dðCuzÞ dt ^ k Since A, B, and C are constant, this becomes _ H rel ¼ A _ ux ^ i þ B _ uy ^ j þ C _ uz ^ k (9.70) Substituting Eqns (9.67) and (9.70) into Eqn (9.68) yields Mnet ¼ A _ ux ^ i þ B _ uy ^ j þ C _ uz ^ k þ ^ i ^ j ^ k Ux Uy Uz Aux Buy Cuz Expanding the crossproduct and collecting the terms leads to Mxnet ¼ A _ ux þ CUyuz BUzuy Mynet ¼ B _ uy þ AUzux CUxuz Mznet ¼ C _ uz þ BUxuy AUyux (9.71) 9.6 Euler’s equations 497
  • 40. If the comoving frame is a rigidly attached body frame, then its angular velocity is the same as that of the body, that is, U ¼ u. In that case, Eqn (9.68) reduces to the classical Euler’s equation of motion, namely, Mnet ¼ _ H rel þ u H (9.72a) the three components of which are obtained from Eqn (9.71) as Mxnet ¼ A _ ux þ ðC BÞuyuz Mynet ¼ B _ uy þ ðA CÞuzux Mznet ¼ C _ uz þ ðB AÞuxuy (9.72b) Equation (9.68) is sometimes referred to as the modified Euler equation. When U ¼ u, it follows from Eqn (9.69) that _ ux ¼ ax _ uy ¼ ay _ uz ¼ az (9.73) That is, the relative angular acceleration equals the absolute angular acceleration when U ¼ u. Rather than calculating the time derivatives _ ux, _ uy, and _ uz for use in Eqn (9.72), we may in this case first compute a in the absolute XYZ frame a ¼ du dt ¼ duX dt ^ I þ duY dt ^ J þ duZ dt ^ K and then project these components onto the xyz body frame, so that 8 : _ ux _ uy _ uz 9 = ; ¼ ½QXx 8 : duX=dt duY=dt duZ=dt 9 = ; (9.74) where [Q]Xx is the time-dependent orthogonal transformation from the inertial XYZ frame to the noninertial xyz frame. EXAMPLE 9.12 Calculate the net moment on the solar panel of Examples 9.2 and 9.8 (Figure 9.17). Solution Since the comoving frame is rigidly attached to the panel, Euler’s equation (Eqn (9.72a)) applies to this problem, that is MGnet ¼ _ HG rel þ u HG (a) where HG ¼ Aux ^ i þ Buy ^ j þ Cuz ^ k (b) and _ HG rel ¼ A _ ux ^ i þ B _ uy ^ j þ C _ uz ^ k (c) 498 CHAPTER 9 Rigid Body Dynamics
  • 41. In Example 9.2, the angular velocity of the panel in the satellite’s x0y0z0 frame was found to be u ¼ _ q^ j 0 þ N^ k 0 (d) In Example 9.8, we showed that the transformation from the panel’s xyz frame to that of the satellite is represented by the matrix ½Q ¼ 2 6 6 4 sin q 0 cos q 0 1 0 cos q 0 sin q 3 7 7 5 (e) We use the transpose of [Q] to transform the components of u into the panel frame of reference, fugxyz ¼ ½QT fugx0y0z0 ¼ 2 6 6 4 sin q 0 cos q 0 1 0 cos q 0 sin q 3 7 7 5 8 : 0 _ q N 9 = ; ¼ 8 : N cos q _ q N sin q 9 = ; or ux ¼ N cos q uy ¼ _ q uz ¼ N sin q (f) In Example 9.2, N and _ q were said to be constant. Therefore, the time derivatives of Eqn (f) are _ ux ¼ dðN cos qÞ dt ¼ N _ q sin q _ uy ¼ d_ q dt ¼ 0 _ uz ¼ dðN sin qÞ dt ¼ N _ q cos q (g) In Example 9.8, the moments of inertia in the panel frame of reference were listed as A ¼ 1 12 m ‘2 þ t2 B ¼ 1 12 m w2 þ t2 C ¼ 1 12 m w2 þ ‘2 IGxy ¼ IGxz ¼ IGyz ¼ 0 (h) θ FIGURE 9.17 Free-body diagram of the solar panel in Examples 9.2 and 9.8. 9.6 Euler’s equations 499
  • 42. Substituting Eqns (b), (c), (f), (g) and (h) into Eqn (a) yields MGnet ¼ 1 12 m ‘2 þ t2 N _ q sin q ^ i þ 1 12 m w2 þ t2 $0$^ j þ 1 12 m w2 þ ‘2 N _ q cos q ^ k þ ^ i ^ j ^ k N cos q _ q N sin q 1 12 m ‘2 þ t2 ðN cos qÞ 1 12 m w2 þ t2 _ q 1 12 m w2 þ ‘2 ðN sin qÞ Upon expanding the cross product and collecting terms, this reduces to MGnet ¼ 1 6 mt2 N _ q sin q^ i þ 1 24 m t2 w2 N2 sin 2q^ j þ 1 6 mw2 N _ q cos q^ k Using the numerical data of Example 9.8 (m ¼ 50 kg, N ¼ 0.1 rad/s, q ¼ 40, _ q ¼ 0:01 rad=s, [ ¼ 6 m, w ¼ 2 m, and t ¼ 0.025 m), we find MGnet ¼ 3:348 106^ i 0:08205^ j þ 0:02554^ k ðN$mÞ EXAMPLE 9.13 Calculate the net moment on the gyro rotor of Examples 9.3 and 9.6. Solution Figure 9.18 is a free-body diagram of the rotor. Since in this case the comoving frame is not rigidly attached to the rotor, we must use Eqn (9.68) to find the net moment about G, that is MGnet ¼ _ HG rel þ U HG (a) where HG ¼ Aux ^ i þ Buy ^ j þ Cuz ^ k (b) and _ HG rel ¼ A _ ux ^ i þ B _ uy ^ j þ C _ uz ^ k (c) From Eqn (i) of Example 9.3, we know that the components of the angular velocity of the rotor in the moving reference frame are ux ¼ _ q uy ¼ N sin q uz ¼ uspin þ N cos q (d) 500 CHAPTER 9 Rigid Body Dynamics
  • 43. Since, as specified in Example 9.3, _ q, N, and uspin are all constant, it follows that _ ux ¼ d_ q dt ¼ 0 _ uy ¼ dðN sin qÞ dt ¼ N _ q cos q _ uz ¼ d uspin þ N cos q dt ¼ N _ q sin q (e) The angular velocity U of the comoving xyz frame is that of the gimbal ring, which equals the angular velocity of the rotor minus its spin. Therefore, Ux ¼ _ q Uy ¼ N sin q Uz ¼ N cos q (f) In Example 9.6, we found that A ¼ B ¼ 1 12 mt2 þ 1 4 mr2 C ¼ 1 2 mr2 (g) FIGURE 9.18 Free-body diagram of the gyro rotor of Examples 9.6 and 9.3. 9.6 Euler’s equations 501
  • 44. Substituting Eqns (b) through (g) into Eqn (a), we get MGnet ¼ 1 12 mt2 þ 1 4 mr2 $0^ i þ 1 12 mt2 þ 1 4 mr2 N _ q cos q ^ j þ 1 2 mr2 N _ q sin q ^ k þ ^ i ^ j ^ k _ q N sin q N cos q 1 12 mt2 þ 1 4 mr2 _ q 1 12 mt2 þ 1 4 mr2 N sin q 1 2 mr2 uspin þ N cos q Expanding the cross product, collecting terms, and simplifying leads to MGnet ¼ 1 2 uspin þ 1 12 3 t2 r2 N cos q mr2 N sin q^ i þ 1 6 t2 r2 N cos q 1 2 uspin mr2 _ q^ j 1 2 N _ q sin qmr2^ k (h) In Example 9.3, the following numerical data were provided: m ¼ 5 kg, r ¼ 0.08 m, t ¼ 0.025 m, N ¼ 2.1 rad/s, q ¼ 60, _ q ¼ 4 rad=s, and uspin ¼ 105 rad/s. For this set of numbers, Eqn (h) becomes MGnet ¼ 0:3203^ i 0:6698^ j 0:1164^ k ðN$mÞ 9.7 Kinetic energy The kinetic energy T of a rigid body is the integral of the kinetic energy 1 2 v2dm of its individual mass elements, T ¼ Z m 1 2 v2 dm ¼ Z m 1 2 v$vdm (9.75) where v is the absolute velocity _ R of the element of mass dm. From Figure 9.8, we infer that _ R ¼ _ RG þ _ r. Furthermore, Eqn (1.52) requires that _ r ¼ u r. Thus, v ¼ vG þ u r, which means v$v ¼ ½vG þ u r$½vG þ u r ¼ vG 2 þ 2vG$ðu rÞ þ ðu rÞ$ðu rÞ We can apply the vector identity introduced in Eqn (1.21), namely A$ðB CÞ ¼ B$ðC AÞ (9.76) to the last term to get v$v ¼ vG 2 þ 2vG$ðu rÞ þ u$½r ðu rÞ Therefore, Eqn (9.75) becomes T ¼ Z m 1 2 vG 2 dm þ vG$ 0 @u Z m rdm 1 A þ 1 2 u$ Z m r ðu rÞdm 502 CHAPTER 9 Rigid Body Dynamics
  • 45. Since r is measured from the center of mass, R m rdm ¼ 0. Recall that, according to Eqn (9.34), Z m r ðu rÞdm ¼ HG It follows that the kinetic energy may be written as T ¼ 1 2 mvG 2 þ 1 2 u$HG (9.77) The second term is the rotational kinetic energy TR, TR ¼ 1 2 u$HG (9.78) If the body is rotating about a point P that is at rest in inertial space, we have from Eqn (9.2) and Figure 9.8 that vG ¼ vP þ u rG=P ¼ 0 þ u rG=P ¼ u rG=P It follows that vG 2 ¼ vG$vG ¼ u rG=P $ u rG=P Making use once again of the vector identity in Eqn (9.76), we find vG 2 ¼ u$ h rG=P u rG=P i ¼ u$ rG=P vG Substituting this into Eqn (9.77) yields T ¼ 1 2 u$ h HG þ rG=P mvG i Equation (9.21) shows that this can be written as T ¼ 1 2 u$HP (9.79) In this case, of course, all of the kinetic energy is rotational. In terms of the components of u and H, whether it is HP or HG, the rotational kinetic energy expression becomes, with the aid of Eqn (9.39), TR ¼ 1 2 uxHx þ uyHy þ uyHz ¼ 1 2 P ux uy uz R 2 6 6 4 Ix Ixy Ixz Ixy Iy Iyz Ixz Iyz Iz 3 7 7 5 8 : ux uy uz 9 = ; Expanding, we obtain TR ¼ 1 2 Ixux 2 þ 1 2 Iyuy 2 þ 1 2 Izuz 2 þ Ixyuxuy þ Ixzuxuz þ Iyzuyuz (9.80) 9.7 Kinetic energy 503
  • 46. Obviously, if the xyz axes are principal axes of inertia, then Eqn (9.80) simplifies considerably, TR ¼ 1 2 Aux 2 þ 1 2 Buy 2 þ 1 2 Cuz 2 (9.81) EXAMPLE 9.14 A satellite in a circular geocentric orbit of 300-km altitude has a mass of 1500 kg, and the moments of inertia relative to a body frame with origin at the center of mass G are ½I ¼ 2 6 6 4 2000 1000 2500 1500 3000 1500 2500 1500 4000 3 7 7 5 kg$m2 If at a given instant the components of angular velocity in this frame of reference are u ¼ 1^ i 0:9^ j þ 1:5^ k ðrad=sÞ calculate the total kinetic energy of the satellite. Solution The speed of the satellite in its circular orbit is v ¼ ffiffiffi m r r ¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 398; 600 6378 þ 300 r ¼ 7:7258 km=s The angular momentum of the satellite is fHGg ¼ ½IGfug ¼ 2 6 6 4 2000 1000 2500 1500 3000 1500 2500 1500 4000 3 7 7 5 8 : 1 0:9 1:5 9 = ; ¼ 8 : 6650 5950 9850 9 = ; kg$m2 =s Therefore, the total kinetic energy is T ¼ 1 2 mvG 2 þ 1 2 u$HG ¼ 1 2 $1500$7725:82 þ 1 2 P 1 0:9 1:5 R 8 : 6650 5950 9850 9 = ; ¼ 44:766 106 þ 13; 390 T ¼ 44:766 MJ Obviously, the kinetic energy is dominated by that due to the orbital motion. 9.8 The spinning top Let us analyze the motion of the simple axisymmetric top in Figure 9.19. It is constrained to rotate about point O. The moving coordinate system is chosen to have its origin at O. The z-axis is aligned with the spin axis of the top (the axis of rotational symmetry). The x-axis is the node line, which passes through O 504 CHAPTER 9 Rigid Body Dynamics
  • 47. and is perpendicular to the plane defined by the inertial Z-axis and the spin axis of the top. The y-axis is then perpendicular to x and z, such that ^ j ¼ ^ k ^ i. By symmetry, the moment of inertia matrix of the top relative to the xyz frame is diagonal, with Ix ¼ Iy ¼ A and Iz ¼ C. From Eqns (9.68) and (9.70), we have M0net ¼ A _ ux ^ i þ A _ uy ^ j þ C _ uz ^ k þ ^ i ^ j ^ k Ux Uy Uz Aux Auy Cuz (9.82) The angular velocity u of the top is the vector sum of the spin rate us and the rates of precession up and nutation un, where up ¼ _ f un ¼ _ q (9.83) Thus, u ¼ un ^ i þ up ^ K þ us ^ k From the geometry, we see that ^ K ¼ sin q^ j þ cos q^ k (9.84) Therefore, relative to the comoving system, u ¼ un ^ i þ up sin q^ j þ us þ up cos q ^ k (9.85) FIGURE 9.19 Simple top rotating about the fixed point O. 9.8 The spinning top 505
  • 48. From Eqn (9.85), we see that ux ¼ un uy ¼ up sin q uz ¼ us þ up cos q (9.86) Computing the time rates of these three expressions yields the components of angular acceleration relative to the xyz frame, given by _ ux ¼ _ un _ uy ¼ _ up sin q þ upun cos q _ uz ¼ _ us þ _ up cos q upun sin q (9.87) The angular velocity U of the xyz system is U ¼ up ^ K þ un ^ i, so that, using Eqn (9.84), U ¼ un ^ i þ up sin q^ j þ up cos q^ k (9.88) From Eqn (9.88), we obtain Ux ¼ un Uy ¼ up sin q Uz ¼ up cos q (9.89) In Figure 9.19, the moment about O is that of the weight vector acting through the center of mass G: M0net ¼ d^ k mg^ K ¼ mgd^ k sin q^ j þ cos q^ k or M0net ¼ mgd sin q^ i (9.90) Substituting Eqns (9.86), (9.87), (9.89), and (9.90) into Eqn (9.82), we get mgd sin q^ i ¼ A _ un ^ i þ A _ up sin q þ upun cos q ^ j þ C _ us þ _ up cos q upun sin q ^ k þ ^ i ^ j ^ k un up sin q up cos q Aun Aup sin q C us þ up cos q (9.91) Let us consider the special case in which q is constant, that is, there is no nutation, so that un ¼ _ un ¼ 0. Then, Eqn (9.91) reduces to mgd sin q^ i ¼ A _ up sin q^ j þ C _ us þ _ up cos q ^ k þ ^ i ^ j ^ k 0 up sin q up cos q 0 Aup sin q C us þ up cos q (9.92) Expanding the determinant yields mgd sin q^ i ¼ A _ up sin q^ j þ C _ us þ _ up cos q ^ k þ h Cupus sin q þ ðC AÞup 2 cos q sin q i ^ i Equating the coefficients of^ i,^ j, and ^ k on each side of the equation and assuming that 0 q 180 leads to mgd ¼ Cupus þ ðC AÞup 2 cos q (9.93a) 506 CHAPTER 9 Rigid Body Dynamics
  • 49. A _ up ¼ 0 (9.93b) C _ us þ _ up cos q ¼ 0 (9.93c) Equation (9.93b) implies _ up ¼ 0, and from Eqn (9.93c), it follows that _ us ¼ 0. Therefore, the rates of spin and precession are both constant. From Eqn (9.93a), we find ðA CÞcos qup 2 Cusup þ mgd ¼ 0 (9.94) If the spin rate is zero, Eqn (9.94) yields up us¼0 ¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi mgd ðC AÞcos q s if ðC AÞ cos q 0 (9.95) In this case, the top rotates about O at this rate, without spinning. If A C (prolate), its symmetry axis must make an angle between 90 and 180 to the vertical; otherwise up is imaginary. On the other hand, if A C (oblate), the angle lies between 0 and 90. Thus, in steady rotation without spin, the top’s axis sweeps out a cone that lies either below the horizontal plane (A C) or above the plane (A C). In the special case where (A C) cos q ¼ 0, Eqn (9.94) yields a steady precession rate that is inversely proportional to the spin rate, up ¼ mgd Cus if ðA CÞ cos q ¼ 0 (9.96) If A ¼ C, this precession apparently occurs irrespective of tilt angle q. If A s C, this rate of precession occurs at q ¼ 90, that is, the spin axis is perpendicular to the precession axis. In general, Eqn (9.94) is a quadratic equation in up, so we can use the quadratic formula to find up ¼ C 2ðA CÞcos q us ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi us 2 4mgdðA CÞcos q C2 r ! (9.97) Thus, for a given spin rate and tilt angle q (q s 90), there are two rates of precession _ f. Observe that if (A C)cos q 0, then up is imaginary when us 2 4mgdðA CÞcos q=C2. Therefore, the minimum spin rate required for steady precession at a constant inclination q is usÞmin ¼ 2 C ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi mgdðA CÞcos q p if ðA CÞcos q 0 (9.98) If (A C) cos q 0, the radical in Eqn (9.97) is real for all us. In this case, as us / 0, up approaches the value given above in Eqn (9.95). 9.8 The spinning top 507
  • 50. EXAMPLE 9.15 Calculate the precession rate up for the top of Figure 9.19 if m ¼ 0.5 kg, A (¼ Ix ¼ Iy) ¼ 12 104 kg m2 , C (¼ Iz) ¼ 4.5 104 kg m2 , and d ¼ 0.05 m. Solution For an inclination of, say, 60, (A C) cos q 0, so that Eqn (9.98) requires usÞmin ¼ 407:01 rpm. Let us choose the spin rate to be us ¼ 1000 rpm ¼ 104.7 rad/sec. Then, from Eqn (9.97), the precession rate as a function of the inclination q is given by either one of the following formulas: up ¼ 31:42 1 þ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1 0:3312 cos q p cos q and up ¼ 31:42 1 ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1 0:3312 cos q p cos q (a) These are plotted in Figure 9.20. Figure 9.21 shows an axisymmetric rotor mounted so that its spin axis (z) remains perpendicular to the precession axis (y). In that case, Eqn (9.85) with q ¼ 90 yields u ¼ up ^ j þ us ^ k (9.99) Likewise, from Eqn (9.88), the angular velocity of the comoving xyz system is U ¼ up ^ j. If we assume that the spin rate and precession rate are constant ðdup=dt ¼ dus=dt ¼ 0Þ, then Eqn (9.68), written for the center of mass G, becomes MGnet ¼ U H ¼ up ^ j Aup ^ j þ Cus ^ k (9.100) where A and C are the moments of inertia of the rotor about the x- and z-axes, respectively. Setting Cus ^ k ¼ Hs, the spin angular momentum, and up ^ j ¼ up, we obtain MGnet ¼ up Hs Hs ¼ Cus ^ k (9.101) Since the center of mass is the reference point, there is no restriction on the motion G for which Eqn (9.101) is valid. Observe that the net gyroscopic moment MGnet exerted on the rotor by its supports is 90 ω 180 –6000 –3000 0 6000 45 50 55 θ (°) 180 90 0 0 3000 p (rpm) ωp (rpm) θ (°) 45 135 45 135 (a) (b) FIGURE 9.20 (a) High-energy precession rate (unlikely to be observed). (b) Low energy precession rate (the one most always seen). 508 CHAPTER 9 Rigid Body Dynamics
  • 51. perpendicular to the plane of the spin and the precession vectors. If a spinning rotor is forced to precess, the gyroscopic moment MGnet develops. Or, if a moment is applied normal to the spin axis of a rotor, it will precess so as to cause the spin axis to turn toward the moment axis. EXAMPLE 9.16 A uniform cylinder of radius r, length L, and mass m spins at a constant angular velocity us. It rests on simple supports (which cannot exert couples), mounted on a platform that rotates at an angular velocity of up. Find the reactions at A and B. Neglect the weight (i.e., calculate the reactions due just to the gyroscopic effects). Solution The net vertical force on the cylinder is zero, so the reactions at each end must be equal and opposite in direction, as shown on the free-body diagram insert in Figure 9.22. Noting that the moment of inertia of a uniform cylinder about it axis of rotational symmetry is 1 2mr2, Eqn (9.101) yields RL^ i ¼ up ^ j 1 2 mr2 us ^ k ¼ 1 2 mr2 up us ^ i so that R ¼ mr2upus 2L ω ω FIGURE 9.22 Illustration of the gyroscopic effect. ω ω FIGURE 9.21 A spinning rotor on a rotating platform. 9.8 The spinning top 509
  • 52. 9.9 Euler angles Three angles are required to specify the orientation of a rigid body relative to an inertial frame. The choice is not unique, but there are two sets in common use: the Euler angles and the yaw, pitch, and roll angles. We will discuss each of them in turn. The reader is urged to review Section 4.5 on orthogonal coordinate transformations and, in particular, the discussion of Euler angle sequences. The three Euler angles f, q, and j shown in Figure 9.23 give the orientation of a body- fixed xyz frame of reference relative to the XYZ inertial frame of reference. The xyz frame is obtained from the XYZ frame by a sequence of rotations through each of the Euler angles in turn. The first rotation is around the Z (¼ z1) axis through the precession angle f. This takes X into x1 and Y into y1. The second rotation is around the x2 (¼ x1) axis through the nutation angle q. This carries y1 and z1 into y2 and z2, respectively. The third and final rotation is around the z (¼ z2) axis through the spin angle j, which takes x2 into x and y2 into y. The matrix [Q]Xx of the transformation from the inertial frame to the body-fixed frame is given by the classical Euler angle sequence (Eqn (4.37)): ½QXx ¼ ½R3ðjÞ½R1ðqÞ½R3ðfÞ (9.102) 2 1 3 3 2 1 X Y x1 x2 , Z z1 , y1 x y y2 z2, z φ φ φ θ ψ ψ ψ θ θ FIGURE 9.23 Classical Euler angle sequence. See also Figure 4.14. 510 CHAPTER 9 Rigid Body Dynamics
  • 53. From Eqns (4.32) and (4.34), we have ½R3ðjÞ ¼ 2 6 6 6 4 cos j sin j 0 sin j cos j 0 0 0 1 3 7 7 7 5 ½R1ðqÞ ¼ 2 6 6 6 4 1 0 0 0 cos q sin q 0 sin q cos q 3 7 7 7 5 ½R3ðfÞ ¼ 2 6 6 6 4 cos f sin f 0 sin f cos f 0 0 0 1 3 7 7 7 5 (9.103) According to Eqn (4.38), the direction cosine matrix is ½QXx ¼ 2 6 6 4 sin f cos q sin j þ cos f cos j cos f cos q sin j þ sin f cos j sin q sin j sin f cos q cos g cos f sin j cos f cos q cos j sin f sin j sin q cos j sin f sin q cos f sin q cos q 3 7 7 5 (9.104) Since this is an orthogonal matrix, the inverse transformation from xyz to XYZ is ½QxX ¼ ð½QXxÞT , ½QxX ¼ 2 6 6 4 sin f cos q sin j þ cos f cos j sin f cos q cos g cos f sin j sin f sin q cos f cos q sin j þ sin f cos j cos f cos q cos j sin f sin j cos f sin q sin q sin j sin q cos j cos q 3 7 7 5 (9.105) Algorithm 4.3 is used to find the three Euler angles q, f, and j from a given direction cosine matrix [Q]Xx. EXAMPLE 9.17 The direction cosine matrix of an orthogonal transformation from XYZ to xyz is ½Q ¼ 2 6 6 4 0:32175 0:89930 0:29620 0:57791 0:061275 0:81380 0:75000 0:43301 0:5000 3 7 7 5 Use Algorithm 4.3 to find the Euler angles f, q, and j for this transformation. Solution Step1 (precession angle): f ¼ tan1 Q31 Q32 ¼ tan1 0:75000 0:43301 ð0 f 360 Þ 9.9 Euler angles 511
  • 54. Since the numerator is negative and the denominator is positive, the angle f lies in the fourth quadrant. f ¼ tan1 ð1:7320Þ ¼ 300 Step 2 (nutation angle): q ¼ cos1 Q33 ¼ cos1 ð0:5000Þ ð0 q 180 Þ q ¼ 120 Step 3 (spin angle): j ¼ tan1Q13 Q23 ¼ tan1 0:29620 0:81380 ð0 j 360 Þ Since both the numerator and denominator are negative, the angle j lies in the third quadrant. j ¼ tan1 ð0:36397Þ ¼ 200 The time rates of change of the Euler angles f, q, and j are, respectively, the precession rate up, the nutation rate un, and the spin us. That is, up ¼ _ f un ¼ _ q us ¼ _ j (9.106) The absolute angular velocity u of a rigid body can be resolved into components ux, uy, and uz along the body-fixed xyz axes, so that u ¼ ux ^ i þ uy ^ j þ uz ^ k (9.107) Figure 9.23 shows that precession is measured around the inertial Z-axis (unit vector ^ K); nutation is measured around the intermediate x1-axis (node line) with unit vector^ i1; and spin is measured around the body-fixed z-axis (unit vector ^ k). Therefore, the absolute angular velocity can alternatively be written in terms of the nonorthogonal Euler angle rates as u ¼ up ^ K þ un ^ i1 þ us ^ k (9.108) In order to find the relationship between the body rates ux, uy, and uz and the Euler angle rates up, un, and us, we must express ^ K and ^ i1 in terms of the unit vectors ^ i^ j^ k of the body-fixed frame. To accomplish that, we proceed as follows. The first rotation ½R3ðfÞ in Eqn (9.102) rotates the unit vectors ^ I^ J^ K of the inertial frame into the unit vectors^ i1 ^ j1 ^ k1 of the intermediate x1y1z1 axes in Figure 9.23. Hence^ i1 ^ j1 ^ k1 are rotated into ^ I^ J^ K by the inverse transformation given by 8 : ^ I ^ J ^ K 9 = ; ¼ ½R3ðfÞT 8 : ^ i1 ^ j1 ^ k1 9 = ; ¼ 2 6 6 4 cos f sin f 0 sin f cos f 0 0 0 1 3 7 7 5 8 : ^ i1 ^ j1 ^ k1 9 = ; (9.109) 512 CHAPTER 9 Rigid Body Dynamics
  • 55. The second rotation [R1(q)] rotates^ i1 ^ j1 ^ k1 into the unit vectors^ i2 ^ j2 ^ k2 of the second intermediate frame x2y2z2 in Figure 9.23. The inverse transformation rotates ^ i2 ^ j2 ^ k2 back into ^ i1 ^ j1 ^ k1: 8 : ^ i1 ^ j1 ^ k1 9 = ; ¼ ½R1ðqÞT 8 : ^ i2 ^ j2 ^ k2 9 = ; ¼ 2 6 6 4 1 0 0 0 cos q sin q 0 sin q cos q 3 7 7 5 8 : ^ i2 ^ j2 ^ k2 9 = ; (9.110) Finally, the third rotation [R3(j)] rotates ^ i2 ^ j2 ^ k2 into ^ i^ j^ k, the target unit vectors of the body-fixed xyz frame. ^ i2 ^ j2 ^ k2 are obtained from ^ i^ j^ k by the reverse rotation, 8 : ^ i2 ^ j2 ^ k2 9 = ; ¼ ½R3ðjÞT 8 : ^ i ^ j ^ k 9 = ; ¼ 2 6 6 4 cos j sin j 0 sin j cos j 0 0 0 1 3 7 7 5 8 : ^ i ^ j ^ k 9 = ; (9.111) From Eqns (9.109), (9.110) and (9.111), we observe that ^ K z}|{ 9:109 ¼ ^ k1 z}|{ 9:110 ¼ sinq^ j2 þ cosq^ k2 z}|{ 9:111 ¼ sinq sinj^ i þ cosj^ j þ cosq^ k or ^ K ¼ sin q sin j^ i þ sin q cos j^ j þ cos q^ k (9.112) Similarly, Eqns (9.110) and (9.111) imply that ^ i1 ¼ ^ i2 ¼ cos j^ i sin j^ j (9.113) Substituting Eqns (9.112) and (9.113) into Eqn (9.108) yields u ¼ up sin q sin j^ i þ sin q cos j^ j þ cos q^ k þ un cos j^ i sin j^ j þ us ^ k or u ¼ up sin q sin j þ un cos j ^ i þ up sin q cos j un sin j ^ j þ us þ up cos q ^ k (9.114) Comparing Eqns (9.107) and (9.114), we see that ux ¼ up sin q sin j þ un cos j uy ¼ up sin q cos j un sin j uz ¼ us þ up cos q (9.115) 9.9 Euler angles 513
  • 56. (Notice that the precession angle f does not appear.) We can solve these three equations to obtain the Euler rates in terms of ux, uy, and uz: up ¼ _ f ¼ 1 sin q ux sin j þ uy cos j un ¼ _ q ¼ ux cos j uy sin j us ¼ _ j ¼ 1 tan q ux sin j þ uy cos j þ uz (9.116) Observe that if ux, uy, and uz are given functions of time, found by solving Euler’s equations of motion (Eqn (9.72)), then Eqns (9.116) are three coupled differential equations that may be solved to obtain the three time-dependent Euler angles, namely f ¼ fðtÞ q ¼ qðtÞ j ¼ jðtÞ With this solution, the orientation of the xyz frame, and hence the body to which it is attached, is known for any given time t. Note, however, that Eqns (9.116) “blow up” when q ¼ 0, that is, when the xy plane is parallel to the XY plane. EXAMPLE 9.18 At a given instant, the unit vectors of a body frame are ^ i ¼ 0:40825^ I 0:40825^ J þ 0:81649^ K ^ j ¼ 0:10102^ I 0:90914^ J 0:40405^ K ^ k ¼ 0:90726^ I þ 0:082479^ J 0:41240^ K (a) and the angular velocity is u ¼ 3:1^ I þ 2:5^ J þ 1:7^ K ðrad=sÞ (b) Calculate up, un, and us (the precession, nutation, and spin rates) at this instant. Solution We will ultimately use Eqn (9.116) to find up, un, and us. To do so we must first obtain the Euler angles f, q, and j as well as the components of the angular velocity in the body frame. The three rows of the direction cosine matrix [Q]Xx comprise the components of the unit vectors ^ i, ^ j, and ^ k, respectively, ½QXx ¼ 2 6 6 4 0:40825 0:40825 0:81649 0:10102 0:90914 0:40405 0:90726 0:082479 0:41240 3 7 7 5 (c) Therefore, the components of the angular velocity in the body frame are fugx ¼ ½QXx fugX ¼ 2 6 6 4 0:40825 0:40825 0:81649 0:10102 0:90914 0:40405 0:90726 0:082479 0:41240 3 7 7 5 8 : 3:1 2:5 1:7 9 = ; ¼ 8 : 0:89817 2:6466 3:3074 9 = ; 514 CHAPTER 9 Rigid Body Dynamics
  • 57. or ux ¼ 0:89817 rad=s uy ¼ 2:6466 rad=s uz ¼ 3:3074 rad=s (d) To obtain the Euler angles f, q, and j from the direction cosine matrix in Eqn (c), we use Algorithm 4.3, as illustrated in Example 9.17. That algorithm is implemented as the MATLAB function dcm_to_Euler.m in Appendix D.20. Typing the following lines in the MATLAB Command Window: Q ¼ [ .40825 -.40825 .81649 -.10102 -.90914 -.40405 .90726 .082479 -.41240]; [phi theta psi] ¼ dcm_to_euler(Q) produces the following output: phi ¼ 95.1945 theta ¼ 114.3557 psi ¼ 116.3291 Substituting q ¼ 114.36 and j ¼ 116.33 together with the angular velocities of Eqn (d) into Eqn (9.116) yields up ¼ 1 sin 114:36 ½ 0:89817$sin 116:33 þ ð2:6466Þ$cos 116:33 ¼ 0:40492 rad=s un ¼ 0:89817$cos 116:33 ð2:6466Þ$sin 116:33 ¼ 2:7704 rad=s us ¼ 1 tan 114:36 ½ 0:89817$sin 116:33 þ ð2:6466Þ$cos 116:33 þ ð3:3074Þ ¼ 3:1404 rad=s EXAMPLE 9.19 The mass moments of inertia of a body about the principal body frame axes with origin at the center of mass G are A ¼ 1000 kg$m2 B ¼ 2000 kg$m2 C ¼ 3000 kg$m2 (a) The Euler angles in radians are given as functions of time in seconds as follows: f ¼ 2te0:05t q ¼ 0:02 þ 0:3 sin 0:25t j ¼ 0:6t (b) At t ¼ 10 s, find (a) The net moment about G and (b) The components aX, aY, and aZ of the absolute angular acceleration in the inertial frame. 9.9 Euler angles 515
  • 58. Solution (a) We must use Euler’s equations (Eqns (9.72)) to calculate the net moment, which means we must first obtain ux, uy, uz, _ ux , _ uy , and _ uz . Since we are given the Euler angles as a function of time, we can compute their time derivatives and then use Eqn (9.115) to find the body frame angular velocity components and their derivatives. Starting with Eqn (b), we get up ¼ df dt ¼ d dt 2te0:05t ¼ 2e0:05t 0:1te0:05t _ up ¼ dup dt ¼ d dt 2e0:05t 0:1e0:05t ¼ 0:2e0:05t þ 0:005te0:05t Proceeding to the remaining two Euler angles leads to un ¼ dq dt ¼ d dt ð0:02 þ 0:3 sin 0:25tÞ ¼ 0:075 cos 0:25t _ un ¼ dun dt ¼ d dt ð0:075 cos 0:25tÞ ¼ 0:01875 sin 0:25t us ¼ dj dt ¼ d dt ð0:6tÞ ¼ 0:6 _ us ¼ dus dt ¼ 0 Evaluating all these quantities, including those in Eqn (b), at t ¼ 10 s yields f ¼ 335:03 up ¼ 0:60653 rad=s _ up ¼ 0:09098 rad=s2 q ¼ 11:433 un ¼ 0:060086 rad=s _ un ¼ 0:011221 rad=s2 j ¼ 343:77 us ¼ 0:6 rad=s _ us ¼ 0 (c) Equation (9.115) relates the Euler angle rates to the angular velocity components, ux ¼ up sin q sin j þ un cos j uy ¼ up sin q cos j un sin j uz ¼ us þ up cos q (d) Taking the time derivative of each of these equations in turn leads to the following three equations: _ ux ¼ upun cos q sin j þ upus sin q cos j unus sin j þ _ up sin q sin j þ _ un cos j _ uy ¼ upun cos q cos j up us sin q sin j unus cos j þ _ up sin q cos j _ un sin j _ uz ¼ upun sin q þ _ up cos q þ _ us (e) Substituting the data in Eqn (c) into Eqns (d) and (e) yields ux ¼ 0:091286 rad=s uy ¼ 0:098649 rad=s uz ¼ 1:1945 rad=s _ ux ¼ 0:063435 rad=s2 _ uy ¼ 2:2346 105 rad=s2 _ uz ¼ 0:08195 rad=s2 (f) With Eqns (a) and (f) we have everything we need for Euler’s equations, namely, Mxnet ¼ A _ ux þ ðC BÞuy uz Mynet ¼ B _ uy þ ðA CÞuz ux Mznet ¼ C _ uz þ ðB AÞux uy 516 CHAPTER 9 Rigid Body Dynamics
  • 59. from which we find Mxnet ¼ 181:27 N$m Mynet ¼ 218:12 N$m Mznet ¼ 254:86 N$m (b) Since the comoving xyz frame is a body frame, rigidly attached to the solid, we know from Eqn (9.74) that 8 : aX aY aZ 9 = ; ¼ ½QxX 8 : _ ux _ uy _ uz 9 = ; (g) In other words, the absolute angular acceleration and the relative angular acceleration of the body are the same. All we have to do is project the components of relative acceleration in Eqn (f) onto the axes of the inertial frame. The required orthogonal transformation matrix is given in Eqn (9.105), ½QxX ¼ 2 6 6 4 sin f cos q sin j þ cos f cos j sin f cos q cos g cos f sin j sin f sin q cos f cos q sin j þ sin f cos j cos f cos q cos j sin f sin j cos f sin q sin q sin j sin q cos j cos q 3 7 7 5 Upon substituting the numerical values of the Euler angles from Eqn (c), this becomes ½QxX ¼ 2 6 6 4 0:75484 0:65055 0:083668 0:65356 0:73523 0:17970 0:055386 0:19033 0:98016 3 7 7 5 Substituting this and the relative angular velocity rates from Eqn (f) into Eqn (g) yields 8 : aX aY aZ 9 = ; ¼ 2 6 6 4 0:75484 0:65055 0:083668 0:65356 0:73523 0:17970 0:055386 0:19033 0:98016 3 7 7 5 8 : 0:063435 2:2345 105 0:08195 9 = ; ¼ 8 : 0:054755 0:026716 0:083833 9 = ; rad=s2 EXAMPLE 9.20 Figure 9.24 shows a rotating platform on which is mounted a rectangular parallelepiped shaft (with dimensions b, h, and [) spinning about the inclined axis DE. If the mass of the shaft is m, and the angular velocities up and us are constant, calculate the bearing forces at D and E as a function of f and j. Neglect gravity, since we are interested only in the gyroscopic forces. (The small extensions shown at each end of the parallelepiped are just for clarity; the distance between the bearings at D and E is [.) Solution The inertial XYZ frame is centered at O on the platform, and it is right handed ð^ I ^ J ¼ ^ KÞ. The origin of the right- handed comoving body frame xyz is at the shaft’s center of mass G, and it is aligned with the symmetry axes of the 9.9 Euler angles 517
  • 60. parallelepiped. The three Euler angles f, q, and j are shown in Figure 9.24. Since q is constant, the nutation rate is zero (un ¼ 0). Thus, Eqn (9.115) reduces to ux ¼ up sin q sin j uy ¼ up sin q cos j uz ¼ up cos q þ us (a) Since up, us, and q are constant, it follows (recalling Eqn (9.106)) that _ ux ¼ up us sin q cos j _ uy ¼ up us sin q sin j _ uz ¼ 0 (b) The principal moments of inertia of the parallelepiped are (Figure 9.10(c)) A ¼ ‘x ¼ 1 12 m h2 þ ‘2 B ¼ ‘y ¼ 1 12 m b2 þ ‘2 C ¼ ‘z ¼ 1 12 m b2 þ h2 (c) Figure 9.25 is a free-body diagram of the shaft. Let us assume that the bearings at D and E are such as to exert just the six body frame components of force shown. Thus, D is a thrust bearing to which the axial torque TD is applied from, say, a motor of some kind. At E, there is a simple journal bearing. D X Y Z y z x O E (Measured in XY plane) (Measured in Zz plane) Shaft Platform ˙ p ˙ s x x (Measured in x'x plane, perpendicular to shaft) xyz axes attached to the shaft l/2 G l/2 b h φ φ θ ψ ψ ω ω = = ′ ′ FIGURE 9.24 Spinning block mounted on a rotating platform. G x y z b h TD Dx Dz Dy Ey Ex l/2 l/2 FIGURE 9.25 Free-body diagram of the block in Figure 9.24. 518 CHAPTER 9 Rigid Body Dynamics