Download as pdf or txt
Download as pdf or txt
You are on page 1of 90

CHAPTER

RIGID BODY DYNAMICS


11
11.1 INTRODUCTION
Just as Chapter 1 provides a foundation for development of 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 of a rigid
body in three dimensions requires taking advantage of the vector nature of angular velocity and know-
ing how to take the time derivative of moving vectors, which is explained in Chapter 1. Several ex-
amples in the current chapter illustrate how this is done.
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 known as the center of mass and applying the methods of particle mechanics to deter-
mine its 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 re-
quires 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 (Ω), the argument of periapsis (ω), and the inclination (i). These
were introduced in Chapter 4 to orient orbits in space. The other set comprises the yaw, pitch, and roll
Orbital Mechanics for Engineering Students. https://doi.org/10.1016/B978-0-08-102133-0.00011-8
# 2020 Elsevier Ltd. All rights reserved.
543
544 CHAPTER 11 RIGID BODY DYNAMICS

angles, which are suitable for describing the orientation of an airplane. Both the Euler angles and the
yaw, pitch, and roll angles will be employed in Chapter 12.
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.

11.2 KINEMATICS
Fig. 11.1 shows a moving rigid body and its instantaneous axis of rotation, which defines the direction
of the absolute angular velocity vector ω. 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 + RB=A

Differentiating this equation through with respect to time, we get


dRB=A
R_ B ¼ R_ A + (11.1)
dt
where R_ A and R_ B 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 Eq. (1.52). That is,
dRB=A
¼ ω  RB=A
dt
Thus, Eq. (11.1) becomes
vB ¼ vA + ω  RB=A (11.2)

Taking the time derivative of Eq. (11.1) yields

€ A + d RB=A
2
€B ¼ R
R (11.3)
dt2

FIG. 11.1
Rigid body and its instantaneous axis of rotation.
11.2 KINEMATICS 545

where R€ A and R
€ B are the absolute accelerations aA and aB of the two points of the rigid body, while from
Eq. (1.53) we have
d2 RB=A  
¼ α  RB=A + ω  ω  RB=A
dt2
where α is the angular acceleration, α ¼ dω=dt. Therefore, Eq. (11.3) can be written
 
aB ¼ aA + α  RB=A + ω  ω  RB=A (11.4)

Eqs. (11.2) and (11.4) 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 ω and angular acceleration α 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 absolute
frame, as illustrated in Fig. 11.2. In that case, the absolute angular velocity ω of body 4 is
ω ¼ ω1 + ω2=1 + ω3=2 + ω4=3 (11.5)

FIG. 11.2
Angular velocity is the vector sum of the relative angular velocities starting with ω1, measured relative to the
inertial frame.
546 CHAPTER 11 RIGID BODY DYNAMICS

Each of these angular velocities is resolved into components along the axes of the moving frame of
reference xyz as shown in Fig. 11.2, so that the vector sum may be written as

ω ¼ ωx^i + ωy^j + ωz k
^ (11.6)

The moving frame is chosen for convenience of analysis, and its own inertial angular velocity vec-
tor is denoted as Ω, as discussed in Section 1.6. According to Eq. (1.56), the absolute angular accel-
eration α ¼ dω=dt is obtained from Eq. (11.6) by means of the following calculation:


α¼ +Ωω (11.7)
dt rel

where


¼ ω_ x^i + ω_ y^j + ω_ z k
^ (11.8)
dt rel

and ω_ x ¼ dωx =dt.


Being able to express the absolute angular velocity vector in an appropriately chosen moving ref-
erence frame, as in Eq. (11.6), is crucial to the analysis of rigid body motion. Once we have the com-
ponents of ω, we simply differentiate each of them with respect to time to arrive at Eq. (11.8). Observe
that the absolute angular acceleration α and dω=dtÞrel , the angular acceleration relative to the moving
frame, are the same if and only if Ω ¼ ω. That occurs if the moving reference is a body-fixed frame (i.e.,
a set of xyz axes imbedded in the rigid body itself).

EXAMPLE 11.1
The airplane in Fig. 11.3 flies at a constant speed v while simultaneously undergoing a constant yaw rate ωyaw about a
vertical axis and describing a circular loop in the vertical plane with a radius ρ. The constant propeller spin rate is ωspin
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 ‘.
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.

FIG. 11.3
Airplane with attached xyz body frame.
11.2 KINEMATICS 547

The absolute angular velocity of the airplane has two components, the yaw rate and the counterclockwise pitch angular
velocity v/ρ of its rotation in the circular loop,

^ ¼ ωyaw^j + v k
ωairplane ¼ ωyaw^j + ωpitch k ^
ρ
The angular velocity of the body-fixed moving frame is that of the airplane, Ω ¼ ωairplane , so that
v^
Ω ¼ ωyaw^j + k (a)
ρ
The absolute angular velocity of the propeller is that of the airplane plus the angular velocity of the propeller relative to the
airplane

ωprop ¼ ωairplane + ωspin^i ¼ Ω + ωspin^i


which means
v^
ωprop ¼ ωspin^i + ωyaw^j + k (b)
ρ
From Eq. (11.2), the velocity of point P on the propeller relative to point H on the hub, vP/H, is given by
vP=H ¼ vP  vH ¼ ωprop  rP=H
where rP/H is the position vector of P relative to H at this instant,

rP=H ¼ ‘^j (c)


Thus, using Eqs. (b) and (c),
   
v^
vP=H ¼ ωspin^i + ωyaw^j + k  ‘^j
ρ
from which
v
vP=H ¼  ‘^i + ωspin ‘k
^
ρ
The absolute angular acceleration of the propeller is found by substituting Eqs. (a) and (b) into Eq. (11.7),

dωprop
αprop ¼ + Ω  ωprop
dt
 rel
    
dωspin ^ dωyaw ^ dðv=ρÞ ^ v^ v^
¼ i+ j+ k + ωyaw^j + k  ωspin^i + ωyaw^j + k
dt dt dt ρ ρ
Since ωspin, ωyaw, v, and ρ are all constant, this reduces to
   
v^ v^
αprop ¼ ωyaw^j + k  ωspin^i + ωyaw^j + k
ρ ρ
Carrying out the cross product yields
v
αprop ¼ ωspin^j  ωyaw ωspin k
^ (d)
ρ
From Eq. (11.4), the acceleration of P relative to H, aP/H, is given by
 
aP=H ¼ aP  aH ¼ αprop  rP=H + ωprop  ωprop  rP=H
Substituting Eqs. (b), (c), and (d) into this expression yields
         
v ^  ‘^j + ωspin^i + ωyaw^j + v k
^  ωspin^i + ωyaw^j + v k
aP=H ¼ ωspin^j  ωyaw ωspin k ^  ‘^j
ρ ρ ρ
548 CHAPTER 11 RIGID BODY DYNAMICS

From this we find that


   v^
 
v
aP=H ¼ ωyaw ωspin ‘^i + ωspin^i + ωyaw^j + k   ‘^i + ωspin ‘k
^
ρ ρ
    2
v

v ^
¼ ωyaw ωspin ‘^i + ωyaw ωspin ‘^i  2 + ωspin 2 ‘^j + ωyaw ‘k
ρ ρ
so that finally,
 
v2 v ^
aP=H ¼ 2ωyaw ωspin ‘^i  + ωspin
2
‘^j + ωyaw ‘k
ρ2 ρ

EXAMPLE 11.2
The satellite in Fig. 11.4 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 θ_ 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.
Solution
(a) Since the moving xyz frame is attached to the body of the spacecraft, its angular velocity is
^
Ω ¼ Nk (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,

ωpanel ¼ θ_ ^j + N k
^ (b)
The position vector of A relative to O is
w w
rA=O ¼  sin θ^i + d^j + cosθk
^ (c)
2 2
According to Eq. (11.2), the velocity of A relative to O is
^i ^j ^
k
vA=O ¼ vA  vO ¼ ωpanel  rA=O ¼ 0 θ_ N
w w
 sin θ d cosθ
2 2

FIG. 11.4
Rotating solar panel on a rotating satellite.
11.2 KINEMATICS 549

from which we get


w  w w
vA=O ¼  θ_ cosθ + Nd ^i  N sin θ^j  θ_ sinθk ^
2 2 2
(b) The absolute angular acceleration of the panel is found by substituting Eqs. (a) and (b) into Eq. (11.7),

dωpanel
αpanel ¼ + Ω  ωpanel
dt
"   rel #
d θ_ ^ dN ^  ^  _ ^ ^

¼ j+ k + N k  θ j + N k
dt dt
Since N and θ_ are constants, this reduces to
_ ^i
αpanel ¼ θN (d)
To find the acceleration of A relative to O, we substitute Eqs. (b) through (d) into Eq. (11.4),
 
aA=O ¼ aA  aO ¼ αpanel  rA=O + ωpanel  ωpanel  rA=O
^i ^j ^
k ^i ^j ^
k
 
¼ _
θN 0 0 + θ_ ^j + N k
^  0 θ_ N
w w w w
 sinθ d cosθ  sin θ d cosθ
2 2 2 2
^i ^j ^
k
 w 
_ ^ _
¼  N θ cosθj + N θd k + ^ 0 θ _ N
2 w w w
 θ_ cosθ  Nd N sinθ  θ_ sinθ
2 2 2
which leads to
w  2 _ 2   w 2
aA=O ¼ N + θ sin θ^i  N Nd + wθ_ cosθ ^j  θ_ cosθk ^
2 2

EXAMPLE 11.3
The gyro rotor illustrated in Fig. 11.5 has a constant spin rate ωspin around axis b–a in the direction shown. The XYZ axes are
fixed. The xyz axes are attached to the gimbal ring, whose angle θ with the vertical is increasing at a constant rate θ_ 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 Fig. 11.6 can be seen to be

^I ¼ cosθ^j + sinθk
^
J^ ¼ ^i (a)
^ ¼ sin θ^j + cosθk
K ^
so that the matrix of the transformation from xyz to XYZ is (Section 4.5)
2 3
0 cosθ sin θ
½QxX ¼ 4 1 0 0 5 (b)
0 sinθ cosθ
550 CHAPTER 11 RIGID BODY DYNAMICS

FIG. 11.5
Rotating, precessing, nutating gyro.

FIG. 11.6
Orientation of the fixed XZ axes relative to the rotating xz axes.
11.2 KINEMATICS 551

The absolute angular velocity of the gimbal ring is that of the base (N K) ^ plus the angular velocity of the gimbal
relative to the base (θ_ ^i), so that
 
^ + θ_ ^i ¼ N sin θ^j + cosθk
ωgimbal ¼ N K ^ + θ_ ^i ¼ θ_ ^i + N sin θ^j + N cosθk
^ (c)

where we made use of Eq. (a)3 above. Since the moving xyz frame is attached to the gimbal, Ω ¼ ωgimbal , so that

Ω ¼ θ_ ^i + N sinθ^j + N cosθk
^ (d)
The absolute angular velocity of the rotor is its spin relative to the gimbal, plus the angular velocity of the gimbal,
^
ωrotor ¼ ωgimbal + ωspin k (e)
From Eq. (c), it follows that
 
ωrotor ¼ θ_ ^i + N sinθ^j + N cosθ + ωspin k^ (f)

Because ^i, ^j, and k


^ move with the gimbal, expression (f) is valid for any time, not just the instant shown in Fig. 11.5.
Alternatively, applying the vector transformation
fωrotor gXYZ ¼ ½QxX fωrotor gxyz (g)
we obtain the angular velocity of the rotor in the inertial frame, but only at the instant shown in the figure (i.e., when the
x axis aligns with the Y axis):
8 9 2 38 9 8 9
< ωX = 0 cosθ sinθ < θ_ = < ωspin sin θ =
ω 4
¼ 1 0 0 5 N sin θ ¼ θ_
: Y; : ; : ;
ωZ 0 sin θ cosθ N cosθ + ωspin N + ωspin cosθ
or
 
ωrotor ¼ ωspin sin θ^I + θ_ ^J + N + ωspin cosθ K^ (h)
_ and
(b) The angular acceleration of the rotor can be found by substituting Eqs. (d) and (f) into Eq. (11.7), recalling that N, θ,
ωspin are independent of time:

dωrotor
αrotor ¼ + Ω  ωrotor
dt rel

^ ^j ^
"     # i k
d θ_ ^ dðN sin θÞ^ d N cosθ + ωspin ^
¼ i+ j+ _
k + θ N sinθ N cosθ
dt dt dt
_θ N sinθ N cosθ + ωspin
  h i
¼ N θ_ cosθ^j  N θ_ sinθk
^ + Nωspin sin θ^i  ωspin θ_ ^j + ð0Þk^

Upon collecting terms, we get


 
αrotor ¼ Nωspin sin θ^i + θ_ N cosθ  ωspin ^j  N θ_ sin θk
^ (i)

This expression, like Eq. (f), is valid at any time.


The components of αrotor along the XYZ axes are found in the same way as for ωrotor ,
fαrotor gXYZ ¼ ½QxX fαrotor gxyz
which means
8 9 2 38 9 8_ _9
Nωspin sin θ > > θω
< αX >
> = 0 cosθ sin θ >
<   = < spin cosθ  N θ >
=
6 7
αY ¼ 4 1 0 0 5 θ_ N cosθ  ωspin ¼ Nωspin sinθ
>
: ; > >
: >
; : > >
;
αZ 0 sin θ cosθ N θ_ sin θ _ spin sinθ
θω
552 CHAPTER 11 RIGID BODY DYNAMICS

or
 
αrotor ¼ θ_ ωspin cosθ  N ^I + Nωspin sin θ^J  θω
_ spin sin θK
^ (j)

Note carefully that Eq. (j) is not simply the time derivative of Eq. (h). Eqs. (h) and (j) are valid only at the instant that the
xyz and XYZ axes have the alignments shown in Fig. 11.5.

11.3 EQUATIONS OF TRANSLATIONAL MOTION


Fig. 11.7 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. Mo-
lecular microstructure does bear on 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 des-
ignated by capital letters. Thus, the position of the center of mass is RG, defined as
ð
mRG ¼ Rdm (11.9)
m

R is the position of a mass element dm within the continuum. Each element of mass is acted on 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 (Eq. 1.38) is written as

dFnet + df net ¼ dmR (11.10)

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
ð ð ð
dFnet + df net ¼ €
Rdm
m m m

FIG. 11.7
Forces on the mass element dm of a continuous medium.
11.4 EQUATIONS OF ROTATIONAL MOTION 553

Ð
Because the internal forces occur in action–reaction pairs, mdfnet ¼ 0. (External forces on the body are
those without an internal reactant; the reactant lies outside the body and, hence, is outside our purview.)
Thus,
ð
Fnet ¼ €
Rdm (11.11)
m
Ð
where Fnet is the resultant external force on the body, Fnet ¼ mdFnet. From Eq. (11.9),
ð
€ ¼ mR
Rdm €G
m

€ G ¼ aG , the absolute acceleration of the center of mass. Therefore, Eq. (11.11) can be written as
where R
€G
Fnet ¼ mR (11.12)
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 turn our attention to rotational motion around
the center of mass. To simplify things, we 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.

11.4 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 Fig. 11.8 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. Fig. 11.7) is
dMP ¼ r  dFnet + r  df net
where r is the position vector of the mass element dm relative to the point P. Writing the right-hand side
as r  (dFnet + dfnet), substituting Eq. (11.10), and integrating over all the mass elements of the body
yields
ð
MP Þnet ¼ €
r  Rdm (11.13)
m

€ is the absolute acceleration of dm relative to the inertial frame and


where R
ð ð
MP Þnet ¼ r  dFnet + r  df net
m m
Ð
But mr  dfnet ¼ 0 because the internal forces occur in action–reaction pairs. Thus,
ð
MP Þnet ¼ r  dFnet
m

which means the net moment includes only the moment of all the external forces on the body.
554 CHAPTER 11 RIGID BODY DYNAMICS

FIG. 11.8
Position vectors of a mass element in a continuum from several key reference points.

 
From the product rule of calculus, we know that d r  R_ =dt ¼ r  R
€ + r_  R,
_ so that the integrand
in Eq. (11.13) may be written as
 
€ ¼ d r  R_  r_  R_
rR (11.14)
dt
Furthermore, Fig. 11.8 shows that r ¼ R  RP, where RP is the absolute position vector of P. It follows
that
 
r_  R_ ¼ R_  R_ P  R_ ¼ R_ P  R_ (11.15)

Substituting Eq. (11.15) into Eq. (11.14) and then moving that result into Eq. (11.13), yields
ð ð
d _
MP Þnet ¼ r  Rdm + R_ P  _
Rdm (11.16)
dt m m

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
ð
HP ¼ _
r  Rdm (11.17)
m

Observing from Fig. 11.8 that r ¼ rG/P + ρ, we can write Eq. (11.17) as
ð ð ð
 
HP ¼ _
rG=P + ρ  Rdm ¼ rG=P  _
Rdm + _
ρ  Rdm (11.18)
m m m

The last term is the absolute angular momentum relative to the center of mass G,
ð
HG ¼ _
ρ  Rdm (11.19)
m
11.4 EQUATIONS OF ROTATIONAL MOTION 555

Furthermore, by the definition of center of mass (Eq. 11.9),


ð
_
Rdm ¼ mR_ G (11.20)
m

Eqs. (11.19) and (11.20) allow us to write Eq. (11.18) as


HP ¼ HG + rG=P  mvG (11.21)

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, Eq. (11.19) can be cast in a much
more useful form by making the substitution (cf. Fig. 11.8) R ¼ RG + ρ, so that
ð ð ð
 
HG ¼ ρ  R_ G + ρ_ dm ¼ ρ  R_ G dm + _
ρ  ρdm
m m m

In the two integrals on the right, the variable is ρ. R_ G is fixed and can therefore be factored out of the
first integral to obtain
ð  ð
HG ¼ ρdm  R_ G + ρ  ρdm
_
m m
Ð
By definition of the center of mass, mρdm ¼ 0 (the position vector of the center of mass relative to
itself is zero), which means
ð
HG ¼ _
ρ  ρdm (11.22)
m
Ð
Since ρ and ρ_ are the position and velocity vectors relative to the center of mass G, m ρ  ρdm
_ is the
total moment about the center of mass of the linear momentum relative to the center of mass, HG)rel. In
other words,
HG ¼ HG Þrel (11.23)

This is a rather surprising fact, hidden in Eq. (11.19), and is true in general for no other point of
the body.
Another useful angular momentum formula, similar to Eq. (11.21), may be found by substituting
R ¼ RP + r into Eq. (11.17),
ð ð  ð
 
HP ¼ r  R_ P + r_ dm ¼ rdm  R_ P + r  r_ dm (11.24)
m m m

The term on the far right is the net moment of relative linear momentum about P,
ð
HP Þrel ¼ r  r_ dm (11.25)
m
Ð
Also, mrdm ¼ mrG/P, where rG/P is the position vector of the center of mass relative to P. Thus,
Eq. (11.24) can be written as
HP ¼ HP Þrel + rG=P  mvP (11.26)
556 CHAPTER 11 RIGID BODY DYNAMICS

Finally, substituting this into Eq. (11.21), solving for HP)rel, and noting that vG  vP ¼ vG/P, yields
HP Þrel ¼ HG + rG=P  mvG=P (11.27)

This expression is useful when the absolute velocity vG of the center of mass, which is required in
Eq. (11.21), is not available.
So far, we have written down some formulas for calculating the angular momentum about an ar-
bitrary 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 Eqs. (11.17) and (11.20) into
Eq. (11.16), we obtain
_ P + R_ P  mR_ G
MP Þnet ¼ H
Thus, for an arbitrary point P,
_ P + vP  mvG
MP Þnet ¼ H (11.28)
where vP and vG are the absolute velocities of points P and G, respectively. This expression is appli-
cable to two important special cases.
If the point P is at rest in inertial space (vP ¼ 0), then Eq. (11.28) reduces to
_P
MP Þnet ¼ H (11.29)
This equation holds as well if vP and vG are parallel (e.g., if P is the point of contact of a wheel rolling
while slipping in the plane). Note that the validity of Eq. (11.29) depends neither on the body being
rigid nor on it being in pure rotation about P. If point P is chosen to be the center of mass, then, since
vG  vG ¼ 0, Eq. (11.28) becomes
_G
MG Þnet ¼ H (11.30)

This equation is valid for any state of motion.


If Eq. (11.30) is integrated over a time interval, then we obtain the angular impulse–momentum
principle,
ð t2
MG Þnet dt ¼ HG Þ2  HG Þ1 (11.31)
t1
Ð
A similar expression follows from Eq. (11.29). Mdt is the angular impulse. If the net angular impulse
is zero, then ΔH 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 Eqs. (11.29) and (11.30) can be obtained that may prove useful in
special circumstances. For example, substituting the expression for HP (Eq. 11.21) into Eq. (11.28)
yields

 
MP Þnet ¼ H_ G + d rG=P  mvG + vP  mvG
dt

_ G+ d
¼H ½ðrG  rP Þ  mvG  + vP  mvG
dt
_ G + ðvG  vP Þ  mvG + rG=P  maG + vP  mvG
¼H
11.5 MOMENTS OF INERTIA 557

or, finally,
_ G + rG=P  maG
MP Þnet ¼ H (11.32)

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 Eq. (11.27) we get
¼0
 zfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflffl{
_P
H _ G + vG=P  mvG=P + rG=P  maG=P
¼H
rel

_ G , invoking Eq. (11.30), and using the fact that aP/G ¼  aG/P leads to
Solving for H

_P
MG Þnet ¼ H + rG=P  maP=G (11.33)
rel

Finally, if the body is rigid, the magnitude of the position vector ρ of any point relative to the center
_
of mass does not change with time. Therefore, Eq. (1.52) requires that ρ5ω  ρ, leading us to conclude
from Eq. (11.22) that
ð
HG ¼ ρ  ðω  ρÞdm ðRigid bodyÞ (11.34)
m

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.
No such simplification of Eq. (11.17) 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 Eq. (1.52) that r_ ¼ ω  r. According to Fig. 11.8,
R ¼ RP + r
Differentiating with respect to time gives
R_ ¼ R_ P + r_ ¼ 0 + ω  r ¼ ω  r
Substituting this into Eq. (11.17) yields the formula for angular momentum in this special case as
ð
HP ¼ r  ðω  rÞdm ðRigid body rotating about fixed point PÞ (11.35)
m

Although Eqs. (11.34) and (11.35) are mathematically identical, we must keep in mind the notation
of Fig. 11.8. Eq. (11.35) applies only if the rigid body is in pure rotation about a stationary point in
inertial space, whereas Eq. (11.34) applies unconditionally to any situation.

11.5 MOMENTS OF INERTIA


To use Eq. (11.29) or Eq. (11.30) 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 focus
on the formula for angular momentum of a rigid body about its center of mass (Eq. (11.34)) because the
expression for fixed point rotation (Eq. (11.35)) is mathematically the same. The integrand of
Eq. (11.34) can be rewritten using the bac–cab vector identity presented in Eq. (2.33),
ρ  ðω  ρÞ ¼ ωρ2  ρðω  ρÞ (11.36)
558 CHAPTER 11 RIGID BODY DYNAMICS

FIG. 11.9
Comoving xyz frame used to compute the moments of inertia.

Let the origin of a comoving xyz coordinate system be attached to the center of mass G, as shown in
Fig. 11.9. The unit vectors of this frame are ^i, ^j, and k.
^ The vectors ρ and ω can be resolved into com-
ponents in the xyz directions to get ρ ¼ xi + yj + zk and ω ¼ ωx^i + ωy^j + ωz k.
^ ^ ^ ^ Substituting these vector
expressions into the right-hand side of Eq. (11.36) yields
     
ρ  ðω  ρÞ ¼ ωx^i + ωy^j + ωz k
^ x2 + y2 + z2  x^i + y^j + zk
^ ωx x + ωy y + ωz z

Expanding the right-hand side and collecting the terms having the unit vectors ^i, ^j, and k
^ in common,
we get
   
ρ  ðω  ρ Þ ¼ y2 + z2 ωx  xyωy  xzωz ^i + yxωx + x2 + z2 ωy  yzωz ^j
  (11.37)
+ zxωx  zyωy + x2 + y2 ωz k ^

We put this result into the integrand of Eq. (11.34) to obtain


HG ¼ Hx^i + Hy^j + Hz k
^ (11.38)

where
8 9 2 38 9
< Hx = Ix Ixy Ixz < ωx =
H ¼ 4 Iyx Iy Iyz 5 ωy (11.39a)
: y; : ;
Hz Izx Izy Iz ωz

or, in matrix notation,


fHg ¼ ½Ifωg (11.39b)

The nine components of the moment of inertia matrix [I] about the center of mass are
ð ð ð
Ix ¼ ðy2 + z2 Þdm Ixy ¼  xydm Ixz ¼  xzdm
mð ð m ðm
Iyx ¼  yxdm Iy ¼ ðx + z Þdm Iyz ¼  yzdm
2 2
(11.40)
ðm mð ð m
Izx ¼  zxdm Izy ¼  zydm Iz ¼ ðx2 + y2 Þdm
m m m
11.5 MOMENTS OF INERTIA 559

Since Iyx ¼ Ixy, Izx ¼ Ixz, and Izy ¼ Iyz, it follows that [I] is a symmetric matrix (i.e., [I]T ¼ [I]). There-
fore, [I] has just 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 Eqs. (11.38) and (11.39) 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 (i.e., its resistance to being set into rotary motion or stopped once rotation is under
way). 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 + z 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,
2 3
A 0 0
½ I ¼ 4 0 B 0 5 (11.41)
0 0 C
where A, B, and C are the principal moments of inertia (all positive), and the xyz axes are the body’s
principal axes of inertia or principal directions. In this case, relative to either the center of mass or a
fixed point of rotation, as appropriate, we have
Hx ¼ Aωx Hy ¼ Bωy Hz ¼ Cωz (11.42)
In general, the angular velocity vector ω and the angular momentum vector H are not parallel
ðω  H 6¼ 0Þ. However, if, for example, ω ¼ ω^i, then according to Eq. (11.42), H ¼ Aω. In other words,
if the angular velocity is aligned with a principal direction, so is the angular momentum. In that case,
the two vectors ω and H are indeed parallel.
Each of the three principal moments of inertia can be expressed as follows:

A ¼ mkx 2 B ¼ mky 2 C ¼ mkz 2 (11.43)


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 Fig. 11.10. 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 Eq. (11.40) 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
2 3
h i mðy2 + z2 Þ mxy mxz
IðmÞ ¼ 4 mxy mðx2 + z2 Þ myz 5 (11.44)
mxz myz mðx2 + y2 Þ
560 CHAPTER 11 RIGID BODY DYNAMICS

FIG. 11.10
Moments of inertia for three common homogeneous solids of mass m. (a) Solid circular cylinder. (b) Circular
cylindrical shell. (c) Rectangular parallelepiped.

EXAMPLE 11.4
The following table lists the mass and coordinates of seven point masses. Find the center of mass of the system and the
moments of inertia about the origin.

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

Solution
The total mass of this system is
X
7
m¼ mi ¼ 28 kg
i¼1

For concentrated masses, the integral in Eq. (11.9) is replaced by the mass times its position vector. Therefore, in this case,
P P
the three components
P7 of the position vector of the center of mass are xG ¼ (1/m) 7i¼1mixi, yG ¼ (1/m) 7i¼1miyi, and
zG ¼ (1/m) i¼1mizi, so that

xG ¼ 0:35 m yG ¼ 0:01964 m zG ¼ 0:4411 m


11.5 MOMENTS OF INERTIA 561

The total moment of inertia is the sum over all the particles of Eq. (11.44) evaluated at each point. Thus,
ð1Þ ð2Þ ð3Þ
zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{
2 3 zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{
2 3 zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{
2 3
0:39 0:3 0:45 5:0575 1:05 0:56 7:25 4 4:5
6 7 6 7 6 7
½I ¼ 4 0:3 1:02 0:18 5 + 4 1:05 1:4 2:1 5+4 4 9:05 3:6 5
0:45 0:18 0:87 0:56 2:1 4:2175 4:5 3:6 8:2
ð4Þ ð5Þ ð6Þ
zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl
2 ffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl3{ zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{
2 3 zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{
2 3
19:515 9:36 9 5:2 3:64 2:08 9:54 1:62 0:9
6 7 6 7 6 7
+ 4 9:36 18:015 9:75 5 + 4 3:64 4:66 2:24 5 + 4 1:62 2:61 4:05 5
9 9:75 18:78 2:08 2:24 7:3 0:9 4:05 7:65
ð7Þ
zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{
2 3
3:6125 2:55 1:275
6 7
+ 4 2:55 2:9725 1:445 5
1:275 1:445 5:14
or
2 3
50:56 20:42 14:94
½I ¼ 4 20:42 39:73 14:90 5ðkg  m2 Þ
14:94 14:90 52:16

EXAMPLE 11.5
Calculate the moments of inertia of a slender, homogeneous straight rod of length ‘ and mass m shown in Fig. 11.11. 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 ρ is uniform and given by
m
ρ¼ (a)

The length of the rod is
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
‘ ¼ a2 + b2 + c2

FIG. 11.11
Uniform slender bar of mass m and length ‘.
562 CHAPTER 11 RIGID BODY DYNAMICS

Starting with Ix we have from Eq. (11.40),


ð‘
 2 2
Ix ¼ y + z ρds
0
in which we replaced the element of mass dm by ρds, 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:
s s s
x¼ a y¼ b z¼ c
‘ ‘ ‘
Thus,
ð‘   ð
s 2 s 2 b2 + c2 ‘ 2 1  
Ix ¼ b + c ρds ¼ ρ 2 s ds ¼ ρ b2 + c2 ‘
0 ‘2
‘2
‘ 0 3
Substituting Eq. (a) yields

1  
Ix ¼ m b2 + c2
3
In precisely the same way, we find

1   1  
Iy ¼ m a2 + c2 Iz ¼ m a2 + b2
3 3
For Ixy we have
ð‘ ð‘  ð
s s  ab ‘ 1
Ixy ¼  xyρds ¼  a b ρds ¼ ρ 2 s2 ds ¼  ρab‘
0 0 ‘ ‘ ‘ 0 3
Once again using Eq. (a),

1
Ixy ¼  mab
3
Likewise,

1 1
Ixz ¼  mac Iyz ¼  mbc
3 3

EXAMPLE 11.6
The gyro rotor (Fig. 11.12) in Example 11.3 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, θ_ ¼ 4 rad=s, ω ¼ 10.5 rad/s, and θ ¼ 60°, calculate
(a) the angular momentum of the rotor about its center of mass G in the body-fixed xyz frame and
(b) the angle between the rotor’s angular velocity vector and its angular momentum vector.
Solution
Eq. (f) from Example 11.3 gives the components of the absolute angular velocity of the rotor in the moving xyz frame.

ωx ¼ θ_ ¼ 4 rad=s
ωy ¼ N sin θ ¼ 2:1: sin 60° ¼ 1:819 rad=s (a)
ωz ¼ ωspin + N cosθ ¼ 10:5 + 2:1: cos60° ¼ 11:55 rad=s
Therefore,

ω ¼ 4^i + 1:819^j + 11:55k


^ ðrad=sÞ (b)
11.5 MOMENTS OF INERTIA 563

FIG. 11.12
Rotor of the gyroscope in Fig. 11.4.

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, Ixy ¼ Ixz ¼ Iyz ¼ 0.
From Fig. 11.10A, we see that the nonzero diagonal entries in the moment of inertia tensor are
1 2 1 2 1 1
A¼B¼ mt + mr ¼ ð5Þð0:025Þ2 + ð5Þð0:08Þ2 ¼ 0:008260 kg  m2
12 4 12 4 (c)
1 2 1 2
C ¼ mr ¼ ð5Þð0:08Þ ¼ 0:0160 kg  m2
2 2
We can use Eq. (11.42) 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 Eq. 11.42).
Substituting Eqs. (a) and (c) into Eq. (11.42) yields
Hx ¼ Aωx ¼ ð0:008260Þð4Þ ¼ 0:03304 kg  m2 =s
Hy ¼ Bωy ¼ ð0:008260Þð1:819Þ ¼ 0:0150 kg  m2 =s (d)
Hz ¼ Cωz ¼ ð0:0160Þð11:55Þ ¼ 0:1848 kg  m2 =s
so that
H ¼ 0:03304^i + 0:0150^j + 0:1848k
^ ðkg  m2 =sÞ (e)
The angle ϕ between H and ω is found by taking the dot product of the two vectors,
   
Hω 2:294
ϕ ¼ cos 1 ¼ cos 1 ¼ 9:717° (f)
Hω 0:1883  12:36
As this problem illustrates, the angular momentum and the angular velocity are in general not collinear.

Consider a Cartesian coordinate system x0 y0 z0 with the same origin as xyz but a different orientation. Let
[Q] be the orthogonal matrix ([Q]1 ¼ [Q]T) that transforms the components of a vector from the xyz
system to the x0 y0 z0 frame. Recall from Section 4.5 that the rows of [Q] are the direction cosines of the
x0 y0 z0 axes relative to xyz. If {H0 } comprises the components of the angular momentum vector along the
x0 y0 z0 axes, then {H0 } is obtained from its components {H} in the xyz frame by the relation

fH0 g ¼ ½QfHg
564 CHAPTER 11 RIGID BODY DYNAMICS

From Eq. (11.39), we can write this as


fH0 g ¼ ½Q½Ifωg (11.45)
where [I] is the moment of inertia matrix (Eqs. 11.39a and 11.39b) in xyz coordinates. Like the angular
momentum vector, the components fωg of the angular velocity vector in the xyz system are related to
those in the primed system (fω0 g) by the expression
fω0 g ¼ ½Qfωg
The inverse relation is simply
fωg ¼ ½Q1 fω0 g ¼ ½QT fω0 g (11.46)
Substituting this into Eq. (11.45), we get
fH0 g ¼ ½Q½I½QT fω0 g (11.47)
But the components of angular momentum and angular velocity in the x0 y0 z0 frame are related by an
equation of the same form as Eq. (11.39), so that
fH0 g ¼ ½I0 fω0 g (11.48)
0
where [I ] comprises the components of the inertia matrix in the primed system. Comparing the right-
hand sides of Eqs. (11.47) and (11.48), we conclude that
½I0  ¼ ½Q½I½QT (11.49a)
That is,
2 3 2 32 32 3
Ix0 Ix0 y0 Ix0 z0 Q11 Q12 Q13 Ix Ixy Ixz Q11 Q21 Q31
6 7 6 76 76 7
6 Iy0 x0 Iy0 Iy0 z0 7 ¼ 6 Q21 Q22 Q23 76 Iyx Iy Iyz 76 Q12 Q22 Q32 7 (11.49b)
4 5 4 54 54 5
Iz0 x0 Iz0 y0 Iz0 Q31 Q32 Q33 Izx Izy Iz Q13 Q23 Q33

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,
bRow 1c
T

bRow 1c 2 Ix Ixy 3 zfflfflfflffl}|fflfflfflffl{


8 9
Ixz > Q11 >
zfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflffl{ 6 < =
7
Ix0 ¼ bQ11 Q12 Q13 c 4 Iyx Iy Iyz 5 Q12
>
: >
;
Izx Izy Iz Q13
(11.50)
bRow 3c
T

2 38zfflfflfflffl}|fflfflfflffl{
9
bRow 2c Ix Ixy Ixz > Q31 >
zfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflffl{ 6 < =
7
Iy0 z0 ¼ b Q21 Q22 Q23 c 4 Iyx Iy Iyz 5 Q32
>
: >
;
Izx Izy Iz Q33
etc.
Any object represented by a square matrix whose components transform according to Eq. (11.49) is
called a second-order tensor. We may therefore refer to [I] as the inertia tensor.
11.5 MOMENTS OF INERTIA 565

EXAMPLE 11.7
Find the mass moment of inertia of the system of point masses in Example 11.4 about an axis from the origin through the
point with coordinates (2 m, 3 m, 4 m).
Solution
From Example 11.4, the moment of inertia tensor for the system of point masses is
2 3
50:56 20:42 14:94
6 7 
½I ¼ 6 7
4 20:42 39:73 14:90 5 kg  m
2

14:94 14:90 52:16


The vector V connecting the origin with (2 m, 3 m, 4 m) is

V ¼ 2^i  3^j + 4k
^
The unit vector in the direction of V is
V
^V ¼
u ¼ 0:3714^i  0:5571^j + 0:7428k
^
kVk
^V as the unit vector along the x0 axis of a rotated Cartesian coordinate system. Then, from Eq. (11.50) we get
We may consider u
2 32 3
50:56 20:42 14:94 0:3714
6 76 7
IV 0 ¼ b 0:3714 0:5571 0:7428 c6 76 7
4 20:42 39:73 14:90 54 0:5571 5
14:94 14:90 52:16 0:7428
8 9
>
> 3:695 >
>
< =
¼ b 0:3714 0:5571 0:7428 c 3:482 ¼ 19:06 kg  m2
>
> >
>
: ;
24:90

EXAMPLE 11.8
For the satellite of Example 11.2, which is reproduced in Fig. 11.13, the data are as follows: N ¼ 0.1 rad/s and
θ_ ¼ 0:01 rad=s, in the directions shown. θ 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.

FIG. 11.13
Satellite and solar panel.
566 CHAPTER 11 RIGID BODY DYNAMICS

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 Fig. 11.10C, the moments of inertia of the panel relative to the xyz
coordinate system are
1  2 2 1  
IG Þx ¼ m ‘ + t ¼  50  62 + 0:0252 ¼ 150:0 kg  m2
12 12
1   1  
IG Þy ¼ m w2 + t2 ¼  50  22 + 0:0252 ¼ 16:67 kg  m2
12 12 (a)
1   1  
IG Þz ¼ m w2 + ‘2 ¼  50  22 + 62 ¼ 166:7 kg  m2
12 12
IG Þxy ¼ IG Þxz ¼ IG Þyz ¼ 0
In matrix notation, 2 3
150:0 0 0  
½IG  ¼ 4 0 16:67 0 5 kg  m2 (b)
0 0 166:7
The unit vectors of the satellite’s x0 y0 z0 system are related to those of the panel’s xyz frame by inspection.
^i0 ¼ sin θ^i + cosθk
^ ¼ 0:6428^i + 0:7660k
^
^j0 ¼ ^j (c)
^0 ¼ cosθ^i + sinθk
k ^ ¼ 0:7660^i + 0:6428k
^
0 0 0
The matrix [Q] of the transformation from xyz to x0 y0 z0 comprises the direction cosines of ^i , ^j , and k
^ , which we infer from
Eqs. (c),
2 3
0:6428 0 0:7660
½Q ¼ 4 0 1 0 5 (d)
0:7660 0 0:6428
In Example 11.2, we found that the absolute angular velocity of the panel, in the satellite’s x0 y0 z0 frame of reference, is
0 0 0 0
^ ¼ 0:01^j + 0:1k
ω ¼ θ_ ^j + N k ^ ðrad=sÞ
That is, 8 9
< 0 =
fω0 g ¼ 0:01 ðrad=sÞ (e)
: ;
0:1
To find the absolute angular momentum {H0 G} of the panel in the satellite system requires the use of Eq. (11.39),
 0
HG ¼ I0G fω0 g (f)
Before doing so, we must transform the components of the moments of the inertia tensor in Eq. (b) from the unprimed
(panel) system to the primed (satellite) system, by means of Eq. (11.49),

I0G ¼ ½Q½IG ½QT


2 32 32 3
0:6428 0 0:7660 150 0 0 0:6428 0 0:7660
6 76 76 7
¼4 0 1 0 54 0 16:67 0 54 0 1 0 5
0:7660 0 0:6428 0 0 166:7 0:7660 0 0:6428
so that 2 3
159:8 0 8:205  
I0G ¼ 4 0 16:67 0 5 kg  m2 (g)
8:205 0 156:9
11.5 MOMENTS OF INERTIA 567

Then Eq. (f) yields


2 38 9 8 9
8:205 >
159:8 >00 > > >
> 0:8205 >
>
 0 6 7< = < = 
HG ¼ 6 7
4 0 16:67 0 5> 0:01 > ¼ > 0:1667 > kg  m =s
2
>
: > >
; : >
;
8:205 0 156:9 0:1 15:69
or, in vector notation,
0 0  
HG ¼ 0:8205^i  0:1667^j + 15:69k0 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 Eq. (11.27) to calculate
the angular momentum HO)rel relative to the satellite’s center of mass O,
HO Þrel ¼ HG + rG=O  mvG=O (i)
rG/O is the position vector from O to G,
   
‘ 0 6 0 0
rG=O ¼ dO + ^j ¼ 1:5 + ^j ¼ 4:5^j ðmÞ (j)
2 2
The velocity of G relative to O, vG/O, is found from Eq. (11.2),

vG=O ¼ ωsatellite  rG=O ¼ N k ^0  4:5^j0 ¼ 0:45^i0 ðm=sÞ


^0  rG=O ¼ 0:1k (k)
Substituting Eqs. (h), (j), and (k) into Eq. (i) finally yields
 0 0
 h  i
^0 + 4:5^j0  50 0:45^i0
HO Þrel ¼ 0:8205^i  0:1667^j + 15:69k
(l)
0 0
^0 ðkg  m2 =sÞ
¼ 0:8205^i  0:1667^j + 116:9k

Note that we were unable to use Eq. (11.21) 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 Eq. (11.49) will yield a moment of inertia
matrix [I0 ] that is diagonal (i.e., of the form given by Eq. (11.41))? 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 ω be parallel to the principal direction defined by the vector e, so that
ω ¼ βe, where β is a scalar. Since ω points in the principal direction of the inertia tensor, so must H,
which means H is also parallel to e. Therefore, H ¼ αe, where α is a scalar. From Eq. (11.39), it follows
that
αfeg ¼ fIgðβfegÞ

or
½Ifeg ¼ λfeg

where λ ¼ α/β (a scalar). That is,


2 38 9 8 9
Ix Ixy Ixz >> ex >
> >
> ex >
6 7< = < > =
6 Ixy Iy Iyz 7 ey ¼ λ ey
4 5> > > >
>
: > ; >
: > ;
Ixz Iyz Iz ez ez
568 CHAPTER 11 RIGID BODY DYNAMICS

This can be written


2 38 9 8 9
Ix  λ Ixy Ixz < ex = < 0 =
4 Ixy Iy  λ Iyz 5 ey ¼ 0 (11.51)
: ; : ;
Ixz Iyz Iz  λ ez 0

The trivial solution of Eq. (11.51) is e ¼ 0, which is of no interest. The only way that Eq. (11.51)
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  λ Ixy Ixz
Ixy Iy  λ Iyz ¼ 0 (11.52)
Ixz Iyz Iz  λ

Expanding the determinant, we find


Ix  λ Ixy Ixz
Ixy Iy  λ Iyz ¼ λ3 + J1 λ2  J2 λ + J3 (11.53)
Ixz Iyz Iz  λ

where
Ix Ixy Ixz
Ix Ixy Ix Ixz Iy Iyz
J1 ¼ Ix + Iy + Iz J2 ¼ + + J3 ¼ Ixy Iy Iyz (11.54)
Ixy Iy Ixz Iz Iyz Iz
Ixz Iyz Iz

J1, J2, and J3 are invariants (i.e., they have the same value in every Cartesian coordinate system).
Eqs. (11.52) and (11.53) yield the characteristic equation of the tensor [I],
λ3  J1 λ2 + J2 λ  J3 ¼ 0 (11.55)
The three roots λp (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, λp back
into Eq. (11.51) to obtain
2 38 ðpÞ 9 8 9
Ix  λp Ixy Ixz > ex >
< = <0=
4 Ixy Iy  λp Iyz 5 eðpÞ ¼ 0 ðp ¼ 1, 2, 3Þ (11.56)
> y > : ;
Ixz Iyz Iz  λp : eðzpÞ ; 0

Solving this system yields the three eigenvectors e(p) corresponding to each of the three eigenvalues λp.
The three eigenvectors are orthogonal, also due to the symmetry of matrix [I]. Each eigenvalue is a
principal moment of inertia and its corresponding eigenvector is a principal direction.

EXAMPLE 11.9
Find the principal moments of inertia and the principal axes of inertia of the inertia tensor
2 3
100 20 100
6 7
½I ¼ 4 20 300 50 5kg  m2
100 50 500
11.5 MOMENTS OF INERTIA 569

Solution
We seek the nontrivial solutions of the system [I]{e} ¼ λ{e}. That is,
2 38 9 8 9
100  λ 20 100 < ex = < 0 =
4 20 300  λ 50 5 ey ¼ 0 (a)
: ; : ;
100 50 500  λ ez 0
From Eq. (11.54),
J1 ¼ 100 + 300 + 500 ¼ 900
100 20 100 100 300 50
J2 ¼ + + ¼ 217,100
20 300 100 500 50 500
(b)
100 20 100
J3 ¼ 20 300 50 ¼ 11,350,000
100 50 500
Thus, the characteristic equation is
λ3  900λ2 + 217,100λ  11,350,000 ¼ 0 (c)
The three roots are the principal moments of inertia, which are found to be
 
λ1 ¼ 532:052 λ2 ¼ 295:840 λ3 ¼ 72:1083 kg  m2 (d)
We substitute each of these, in turn, back into Eq. (a) to find its corresponding principal direction.
Substituting λ1 ¼ 532.052 kg  m2 into Eq. (a) we obtain
2 38 ð1Þ 9 8 9
432:052 20:0000 100:0000 > < ex >= > <0> =
6 7
4 20:0000 232:052 50:0000 5 eðy1Þ ¼ 0 (e)
>
: ð1Þ ; : >
> > ;
100:0000 50:0000 32:0519 ez 0
Since the determinant of the coefficient matrix is zero, at most two of the three equations in Eq. (e) are independent. 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
x ¼ 1 and solve for ey and ez using any two of the independent equations in Eq. (e). With ex ¼ 1, the first two of
e(1) (1) (1) (1)

Eq. (e) become

20:0000eðy1Þ  100:000eðz1Þ ¼ 432:052


(f)
232:052eðy1Þ  50:000eðz1Þ ¼ 20:0000

y and ez yields, together with the assumption that ex ¼ 1,


Solving these two equations for e(1) (1) (1)

eðx1Þ ¼ 1:00000 eðy1Þ ¼ 0:882793 eðz1Þ ¼ 4:49708 (g)


The unit vector in the direction of e (1)
is

eð1Þ 1:00000^i + 0:882793^j  4:49708k ^


^e1 ¼ ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ke k
ð 1Þ
1:000002 + 0:8827932 + ð4:49708Þ2
or

^e1 ¼ 0:213186^i + 0:188199^j  0:958714k


^ ðλ1 ¼ 532:052 kg  m2 Þ (h)
Substituting λ2 ¼ 295.840 kg  m into Eq. (a) and proceeding as above we find that
2

^e2 ¼ 0:17632^i  0:972512^j  0:151609k


^ ðλ2 ¼ 295:840 kg  m2 Þ (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.
570 CHAPTER 11 RIGID BODY DYNAMICS

To obtain the third principal direction ^e3 , we can substitute λ3 ¼ 72.1083 kg  m2 into Eq. (a) and proceed 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 Eqs. (h) and (i) into this cross product, we find that

^e3 ¼ 0:960894^i  0:137114^j  0:240587k


^ ðλ3 ¼ 72:1083 kg  m2 Þ (j)
We can check our work by substituting λ3 and ^e3 into Eq. (a) and verify that it is indeed satisfied:
2 38 9 8 9
100  72:1083 20 100 < 0:960894 = <0=
4 20 300  72:1083 50 5 0:137114 verify
¼ 0 (k)
: ; : ;
100 50 500  72:1083 0:240587 0
The components of the vectors ^e1 , ^e2 , and ^e3 define the three rows of the orthogonal transformation [Q] from the xyz
system into the x0 y0 z0 system that is aligned along the three principal directions:
2 3
0:213186 0:188199 0:958714
½Q ¼ 4 0:176732 0:972512 0:151609 5 (l)
0:960894 0:137114 0:240587
Indeed, if we apply the transformation in Eq. (11.49), [I0 ] ¼ [Q][I][Q]T, we find
2 32 3
0:213186 0:188199 0:958714 100 20 100
6 76 7
½I0  ¼ 4 0:176732 0:972512 0:151609 54 20 300 50 5
0:960894 0:137114 0:240587 100 50 500
2 3
0:213186 0:176732 0:960894
6 7
 4 0:188199 0:972512 0:137114 5
0:958714 0:151609 0:240587
2 3
532:052
0 0
6 7 
¼4 0 295:840 0 5 kg  m2
0 0 72:1083

An alternative to the above hand calculations in Example 11.9 is to type the following lines in the
MATLAB Command Window:

I = [ 100 –20 –100


–20 300 –50
–100 –50 500];
[eigenVectors, eigenValues] = eig(I)

Hitting the Enter (or Return) key yields the following output to the Command Window:
eigenVectors =
0.9609 0.1767 –0.2132
0.1371 –0.9725 –0.1882
0.2406 –0.1516 0.9587
eigenValues =
72.1083 0 0
0 295.8398 0
0 0 532.0519
11.5 MOMENTS OF INERTIA 571

Two of the eigenvectors delivered by MATLAB are opposite in direction to those calculated in
Example 11.9. This illustrates the fact that we can determine an eigenvector only to within an arbitrary
scalar factor. To show this, suppose e is an eigenvector of the tensor [I] so that [I]{e} ¼ λ{e}. Mul-
tiplying this equation through by an arbitrary scalar a yields ([I]{e})a ¼ (λ{e})a, or [I]{ae} ¼ λ
{ae}, which means that {ae} is an eigenvector corresponding to the same eigenvalue λ.

11.5.1 PARALLEL AXIS THEOREM


Suppose the rigid body in Fig. 11.14 is in pure rotation about point P. Then, according to Eq. (11.39),
 
HP Þrel ¼ ½IP fωg (11.57)

where [IP] is the moment of inertia tensor about P, given by Eq. (11.40) with
x ¼ xG=P + ξ y ¼ yG=P + η z ¼ zG=P + ζ

On the other hand, we have from Eq. (11.27) that


HP Þrel ¼ HG + rG=P  mvG=P (11.58)

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:
   h i
ðmÞ ðmÞ
rG=P  mvG=P  HP ¼ IP fωg (11.59)
rel

where [I(m)
P ], the moment of inertia of the point mass m about P, is obtained from Eq. (11.44), with
x ¼ xG/P, y ¼ yG/P, and z ¼ zG/P. That is,

FIG. 11.14
The moments of inertia are to be computed at P, given their values at G.
572 CHAPTER 11 RIGID BODY DYNAMICS

2   3
m y2G=P + z2G=P mxG=P yG=P mxG=P zG=P
h i 6   7
ðm Þ 6 7
IP ¼ 6 mxG=P yG=P m x2G=P + z2G=P myG=P zG=P 7 (11.60)
4  5
mxG=P zG=P myG=P zG=P m x2G=P + y2G=P

Of course, Eq. (11.39) require


fHG g ¼ ½IG fωg

Substituting this together with Eqs. (11.57) and (11.59) into Eq. (11.58) yields
h i h i
ðmÞ ðm Þ
½IP fωg ¼ ½IG fωg + IP fωg ¼ IG + IP fωg

From this, we may infer the parallel axis theorem,


ðmÞ
I P ¼ IG + I P (11.61)

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 + m y2G=P + z2G=P IPy ¼ IGy + m y2G=P + x2G=P IPz ¼ IGz + m x2G=P + y2G=P
(11.62)
IPxy ¼ IGxy  mxG=P yG=P IPxz ¼ IGxz  mxG=P zG=P IPyz ¼ IGyz  myG=P zG=P

EXAMPLE 11.10
Find the moments of inertia of the rod in Example 11.5 (Fig. 11.15) about its center of mass G.
Solution
From Example 11.5,
2 3
1  2 2 1 1
63 m b + c  mab  mac 7
6 3 3 7
6 1 1  2 2 1 7
½IA  ¼ 6  mab m a +c  mbc 7
6 3 3 3 7
4 1 1 1  2 2 5
 mac  mbc m a +b
3 3 3
Using Eq. (11.62)1 and noting the coordinates of the center of mass in Fig. 11.15,

FIG. 11.15
Uniform slender rod.
11.5 MOMENTS OF INERTIA 573

"  #
h i 1   b 2 c 2 1  
2 2
IGx ¼ IAx  m ðyG  0Þ + ðzG  0Þ ¼ m b + c  m
2 2
+ ¼ m b2 + c2
3 2 2 12

Eq. (11.62)4 yields


1 a b 1
IGxy ¼ IAxy + mðxG  0ÞðyG  0Þ ¼  mab + m   ¼  mab
3 2 2 12
The remaining four moments of inertia are found in a similar fashion, so that
21   1 1 3
m b2 + c2  mab  mac
6 12 2 12 7
6 7
6 1 1  2 2 1 7
½IG  ¼ 6  mab m a +c  mbc 7 (11.63)
6 12 12 12 7
4 5
1 1 1  2 2
 mac  mbc m a +b
12 12 12

EXAMPLE 11.11
Calculate the principal moments of inertia about the center of mass and the corresponding principal directions for the bent
rod in Fig. 11.16. Its mass is uniformly distributed at 2 kg/m.
Solution
The mass of each of the four slender rod segments is
m1 ¼ 2  0:4 ¼ 0:8kg m2 ¼ 2  0:5 ¼ 1kg m3 ¼ 2  0:3 ¼ 0:6kg m4 ¼ 2  0:2 ¼ 0:4kg (a)
The total mass of the system is
X
4
m¼ mi ¼ 2:8kg (b)
i¼1

The coordinates of each segment’s center of mass are


xG1 ¼ 0 yG 1 ¼ 0 zG1 ¼ 0:2m
xG2 ¼ 0 yG2 ¼ 0:25m zG2 ¼ 0:2m
(c)
xG3 ¼ 0:15m yG3 ¼ 0:5m zG3 ¼ 0
xG4 ¼ 0:3m yG4 ¼ 0:4m zG4 ¼ 0

FIG. 11.16
Bent rod for which the principal moments of inertia are to be determined.
574 CHAPTER 11 RIGID BODY DYNAMICS

If the slender rod in Fig. 11.15 is aligned with, say, the x axis, then a ¼ ‘ and b ¼ c ¼ 0, so that according to Eq. (11.63),
2 3
0 0 0
6 1 2 7
60 m‘ 0 7
½IG  ¼ 6 12 7
4 5
1 2
0 0 m‘
12
That is, the moment of inertia of a slender rod about the axes normal to the rod at its center of mass is m‘2/12, 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 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:
   1  
ð1Þ
Ixð1Þ ¼ IG1 + m1 y2G1 + z2G1 ¼  0:8  0:42 + 0:8 0 + 0:22 ¼ 0:04267 kg  m2
x 12
   1  
ð1Þ ð1Þ
Iy ¼ IG1 + m1 x2G1 + z2G1 ¼  0:8  0:42 + 0:8 0 + 0:22 ¼ 0:04267 kg  m2
y 12
  
ð1Þ
Izð1Þ ¼ IG1 + m1 x2G1 + y2G1 ¼ 0 + 0:8ð0 + 0Þ ¼ 0
z
ð1Þ ð1Þ
Ixy ¼ IG1  m1 xG1 yG1 ¼ 0  0:8ð0Þð0Þ ¼ 0
xy

ð1Þ ð1Þ
Ixz ¼ IG1  m1 xG1 zG1 ¼ 0  0:8ð0Þð0:2Þ ¼ 0
xz
ð1Þ ð 1Þ
Ixz ¼ IG1  m1 yG1 zG1 ¼ 0  0:8ð0Þð0Þ ¼ 0
yz

Rod 2:
   1  
ð2Þ
Ixð2Þ ¼ IG2 + m2 y2G2 + z2G2 ¼  1:0  0:52 + 1:0 0 + 0:252 ¼ 0:08333 kg  m2
 x
  12
ð2Þ ð2Þ
Iy ¼ IG2 + m2 x2G2 + z2G2 ¼ 0 + 1:0ð0 + 0Þ ¼ 0
y
   1  
ð2Þ
Izð2Þ ¼ IG2 + m2 x2G2 + y2G2 ¼  1:0  0:52 + 1:0 0 + 0:52 ¼ 0:08333 kg  m2
z 12
ð2Þ ð2Þ
Ixy ¼ IG2  m2 xG2 yG2 ¼ 0  1:0ð0Þð0:5Þ ¼ 0
xy

ð2Þ ð 2Þ
Ixz ¼ IG2  m2 xG2 zG2 ¼ 0  1:0ð0Þð0Þ ¼ 0
xz
ð2Þ ð2Þ
Iyz ¼ IG2  m2 yG2 zG2 ¼ 0  1:0ð0:5Þð0Þ ¼ 0
yz

Rod 3:
    
ð3Þ
Ixð3Þ ¼ IG3 + m3 y2G3 + z2G3 ¼ 0 + 0:6 0:52 + 0 ¼ 0:15 kg  m2
x
   1  
ð3Þ ð2Þ
Iy ¼ IG3 + m3 x2G3 + z2G3 ¼  0:6  0:32 + 0:6 0:152 + 0 ¼ 0:018 kg  m2
y 12
   1  
ð3Þ
Izð3Þ ¼ IG3 + m3 x2G3 + y2G3 ¼  0:6  0:32 + 0:6 0:152 + 0:52 ¼ 0:1680 kg  m2
z 2
ð3Þ ð3Þ
Ixy ¼ IG3  m3 xG3 yG3 ¼ 0  0:6ð0:15Þð0:5Þ ¼ 0:045 kg  m2
xy

ð3Þ ð3Þ
Ixz ¼ IG3  m3 xG3 zG3 ¼ 0  0:6ð0:15Þð0Þ ¼ 0
xz
ð3Þ ð3Þ
Iyz ¼ IG3  m3 yG3 zG3 ¼ 0  0:6ð0:5Þð0Þ ¼ 0
yz
11.5 MOMENTS OF INERTIA 575

Rod 4:
   1  
ð4Þ
Ixð4Þ ¼ IG4 + m4 y2G4 + z2G4 ¼  0:4  0:22 + 0:4 0:42 + 0 ¼ 0:06533 kg  m2
 x
  12
ð4Þ ð4Þ  
Iy ¼ IG4 + m4 x2G4 + z2G4 ¼ 0 + 0:4 0:32 + 0 ¼ 0:0360 kg  m2
y
   1  
ð4Þ
Izð4Þ ¼ IG4 + m4 x2G4 + y2G4 ¼  0:4  0:22 + 0:4 0:32 + 0:42 ¼ 0:1013 kg  m2
z 12
ð4Þ ð4Þ
Ixy ¼ IG4  m4 xG4 yG4 ¼ 0  0:4ð0:3Þð0:4Þ ¼ 0:0480 kg  m2
xy

ð4Þ ð4Þ
Ixz ¼ IG4  m4 xG4 zG4 ¼ 0  0:4ð0:3Þð0Þ ¼ 0
xz
ð4Þ ð4Þ
Iyz ¼ IG4  m4 yG4 zG4 ¼ 0  0:4ð0:4Þð0Þ ¼ 0
yz

The total moments of inertia for all the four rods about O are
X
4 X
4 X
4
Ix ¼ IxðiÞ ¼ 0:3413 kg  m2 Iy ¼ IyðiÞ ¼ 0:09667 kg  m2 Iz ¼ IzðiÞ ¼ 0:3527 kg  m2
i¼1 i¼1 i¼1
(d)
X
4 X
4 X
4
ðiÞ ðiÞ ði Þ
Ixy ¼ Ixy ¼ 0:0930 kg  m2 Ixz ¼ Ixz ¼0 Iyz ¼ Iyz ¼0
i¼1 i¼1 i¼1

The coordinates of the center of mass of the system of four rods are, from Eqs. (a) through (c),

1X 4
1
xG ¼ mi x G i ¼  0:21 ¼ 0:075m
m i¼1 2:8
1X 4
1
yG ¼ mi y G i ¼  0:71 ¼ 0:2536m (e)
m i¼1 2:8
1X 4
1
zG ¼ mi zGi ¼  0:16 ¼ 0:05714m
m i¼1 2:8
We use the parallel axis theorems to shift the moments of inertia in Eq. (d) to the center of mass G of the system
 
IGx ¼ Ix  m y2G + z2G ¼ 0:3413  0:1892 ¼ 0:1522 kg  m2
 
IGy ¼ Iy  m x2G + z2G ¼ 0:09667  0:02489 ¼ 0:17177 kg  m2
 
IGz ¼ Iz  m x2G + y2G ¼ 0:3527  0:1958 ¼ 0:1569 kg  m2
IGxy ¼ Ixy + mxG yG ¼ 0:093 + 0:05325 ¼ 0:03975 kg  m2
IGxz ¼ Ixz + mxG zG ¼ 0 + 0:012 ¼ 0:012 kg  m2
IGyz ¼ Iyz + myG zG ¼ 0 + 0:04057 ¼ 0:04057 kg  m2
Therefore, the inertia tensor, relative to the center of mass, is
2 3 2 3
IGx IGxy IGxz 0:1522 0:03975 0:012
6I 7 6 7 
½I ¼ 4 Gxy IGy IGyz 5 ¼ 4 0:03975 0:07177 0:04057 5 kg  m2 (f)
IGxz IGyz IGz 0:012 0:04057 0:1569
To find the three principal moments of inertia, we may proceed as in Example 11.9, or simply enter the following lines in
the MATLAB Command Window:
IG = [ 0.1522 –0.03975 0.012
–0.03975 0.07177 0.04057
0.012 0.04057 0.1569];
[eigenVectors, eigenValues] = eig(IG)

to obtain
576 CHAPTER 11 RIGID BODY DYNAMICS

eigenVectors =
0.3469 –0.8482 –0.4003
0.8742 0.1378 0.4656
–0.3397 –0.5115 0.7893
eigenValues =
0.0402 0 0
0 0.1658 0
0 0 0.1747

Hence, the three principal moments of inertia and their principal directions are

λ1 ¼ 0:04023 kg  m2 eð1Þ ¼ 0:3469^i + 0:8742^j  0:3397k


^
λ2 ¼ 0:1658 kg  m2 eð2Þ ¼ 0:8482^i + 0:1378^j  0:5115k ^
λ3 ¼ 0:1747 kg  m2 eð3Þ ¼ 0:4003^i + 0:4656^j + 0:7893k
^

11.6 EULER 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 Eqs. (11.29) and (11.30) that
_
Mnet ¼ H (11.64)

Using a comoving coordinate system, with angular velocity Ω and its origin located at the point (G or
P), the angular momentum has the analytical expression
H ¼ Hx^i + Hy^j + Hz k
^ (11.65)

For simplicity, we shall henceforth assume


ðaÞ the moving xyz axes are the principal axes of inertia; (11.66a)
ðbÞ the moments of inertia relative to xyz are constant in time: (11.66b)

Eqs. (11.42) and (11.66a) imply that


H ¼ Aωx^i + Bωy^j + Cωz k
^ (11.67)

where A, B, and C are the principal moments of inertia.


_ ¼ HÞ
According to Eq. (1.56), the time derivative of H is H _ rel + Ω  H, so that Eq. (11.64) can be
written as
_ rel + Ω  H
Mnet ¼ HÞ (11.68)

Keep in mind that, whereas Ω (the angular velocity of the moving xyz coordinate system) and ω (the
angular velocity of the rigid body itself) are both absolute kinematic quantities, Eq. (11.68) contains
their components as projected onto the axes of the noninertial xyz frame given by
ω ¼ ωx^i + ωy^j + ωz k
^
^ ^
Ω ¼ Ωx i + Ωy j + Ωz k^
11.6 EULER EQUATIONS 577

The absolute angular acceleration α is obtained using Eq. (1.56) as


αrel
zfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflffl{
α ¼ ω_ ¼ ω_ x^i + ω_ y^j + ω_ z k ^ +Ωω

That is,
     
α ¼ ω_ x + Ωy ω_ z  Ωz ωy ^i + ω_ y + Ωz ωx  Ωx ωz ^j + ω_ z + Ωx ωy  Ωy ωx k^ (11.69)

Clearly, it is generally true that


αx 6¼ ω_ x αy 6¼ ω_ y αz 6¼ ω_ z

From Eq. (1.57) and Eq. (11.67),  


_ dðAωx Þ^ d Bωy ^ dðCωz Þ ^
HÞrel ¼ i+ j+ k
dt dt dt
Since A, B, and C are constant, this becomes
_ rel ¼ Aω_ x^i + Bω_ y^j + Cω_ z k
HÞ ^ (11.70)

Substituting Eqs. (11.67) and (11.70) into Eq. (11.68) yields


^i ^j ^
k
Mnet ¼ Aω_ x^i + Bω_ y^j + Cω_ z k
^ + Ωx Ωy Ωz
Aωx Bωy Cωz

Expanding the cross product and collecting the terms leads to


Mx Þnet ¼ Aω_ x + CΩy ωz  BΩz ωy
My net ¼ Bω_ y + AΩz ωx  CΩz ωz (11.71)
Mz Þnet ¼ Cω_ z + BΩx ωy  AΩy ωx

If the comoving frame is a body-fixed frame, then its angular velocity vector is the same as that of the
body (i.e., Ω ¼ ω). In that case, Eq. (11.68) reduces to the classical Euler equation of motion; namely,
_ rel + ω  H
Mnet ¼ HÞ (11.72a)
the three components of which are obtained from Eq. (11.71) as
MxÞnet ¼ Aω_ x + ðC  BÞωy ωz
My net ¼ Bω_ y + ðA  CÞωz ωx (11.72b)
Mz Þnet ¼ Cω_ z + ðB  AÞωx ωy

Eq. (11.68) is sometimes referred to as the modified Euler equation.


When Ω ¼ ω, it follows from Eq. (11.69) that
ω_ x ¼ αx ω_ y ¼ αy ω_ z ¼ αz (11.73)

That is, the relative angular acceleration equals the absolute angular acceleration when Ω ¼ ω. Rather
than calculating the time derivatives ω_ x , ω_ y , and ω_ z for use in Eq. (11.72), we may in this case first
compute α in the absolute XYZ frame
dω dωX ^ dωY ^ dωz ^
α¼ ¼ I+ J+ K
dt dt dt dt
578 CHAPTER 11 RIGID BODY DYNAMICS

and then project these components onto the xyz body frame, so that
8 9 8 9
< ω_ x = < dωX =dt =
ω_ ¼ ½QXx dωY =dt (11.74)
: y; : ;
ω_ z dωZ =dt

where [Q]Xx is the time-dependent orthogonal transformation from the inertial XYZ frame to the non-
inertial xyz frame.

EXAMPLE 11.12
Calculate the net moment on the solar panel of Examples 11.2 and 11.8 (Fig. 11.17).
Solution
Since the comoving frame is rigidly attached to the panel, the Euler equation (Eq. 11.72a) applies to this problem. That is
_ G Þrel + ω  HG
MG Þnet ¼ H (a)
where

HG ¼ Aωx^i + Bωy^j + Cωz k


^ (b)
and

_G
H rel
¼ Aω_ x^i + Bω_ y^j + Cω_ z k
^ (c)
0 0 0
In Example 11.2, the angular velocity of the panel in the satellite’s x y z frame was found to be
0
^0
ω ¼ θ_ ^j + N k (d)
In Example 11.8, we showed that the direction cosine matrix for the transformation from the panel’s xyz frame to that of the
satellite is 2 3
sin θ 0 cosθ
½Q ¼ 4 0 1 0 5 (e)
cosθ 0 sinθ
We use the transpose of [Q] to transform the components of ω into the panel frame of reference,
2 38 9 8 9
sinθ 0 cosθ < 0 = < N cosθ =
fωgxyz ¼ ½Q fωgx0 y0 z0 ¼ 4 0
T
1 0 5 θ ¼ _ _
θ
: ; : ;
cosθ 0 sin θ N N sinθ

FIG. 11.17
Free body diagram of the solar panel in Examples 11.2 and 11.8.
11.6 EULER EQUATIONS 579

or

ωx ¼ N cosθ ωy ¼ θ_ ωz ¼ N sin θ (f)


In Example 11.2, N and θ_ were said to be constant. Therefore, the time derivatives of Eq. (f) are

dðN cosθÞ dθ_ dðN sinθÞ


ω_ x ¼ ¼ N θ_ sin θ ω_ x ¼ ¼0 ω_ z ¼ ¼ N θ_ cosθ (g)
dt dt dt
In Example 11.8, the moments of inertia in the panel frame of reference were listed as
1  2 2 1  2 2 1  2 2
A¼ m ‘ +t B¼ m w +t C¼ m w +‘ IG Þxy ¼ IG Þxz ¼ IG Þyz ¼ 0 (h)
12 12 12
Substituting Eqs. (b), (c), (f), (g), and (h) into Eq. (a) yields

1  2 2   1   1   
MG Þnet ¼ m ‘ + t N θ_ sin θ ^i + m w2 + t2 ð0Þ^j + m w2 + ‘2 N θ_ cosθ k^
12 12 12
^i ^j ^
k

+ N cosθ θ_ N sin θ
1  2 2 1  2 2 _ 1  2 2
m ‘ + t ðN cosθÞ m w +t θ m w + ‘ ðN sin θÞ
12 12 12
Upon expanding the cross product determinant and collecting terms, this reduces to

1 1   1
MG Þnet ¼  mt2 N θ_ sinθ^i + m t2  w2 N 2 sin 2θ^j + mw2 N θ_ cosθk
^
6 24 6
Using the numerical data of Example 11.8 (m ¼ 50 kg, N ¼ 0.1 rad/s, θ ¼ 40°, θ_ ¼ 0:01 rad=s, ‘ ¼ 6 m, w ¼ 2 m, and
t ¼ 0.025 m), we find
 
MG Þ ¼ 3:348 106 ^i  0:08205^j + 0:02554k
net
^ðN  mÞ

EXAMPLE 11.13
Calculate the net moment on the gyro rotor of Examples 11.3 and 11.6.
Solution
Fig. 11.18 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 Eq. (11.68) to find the net moment about G. That is
_ G Þrel + Ω  HG
MG Þnet ¼ H (a)
where

HG ¼ Aωx^i + Bωy^j + Cωz k


^ (b)
and
_ G Þrel ¼ Aω_ x i_ + Bω_ y^j + Cω_ z k
H ^ (c)
From Eq. (f) of Example 11.3, we know that the components of the angular velocity of the rotor in the moving reference
frame are

ωx ¼ θ_ ωy ¼ N sin θ ωz ¼ ωspin + N cosθ (d)


580 CHAPTER 11 RIGID BODY DYNAMICS

FIG. 11.18
Free body diagram of the gyro rotor of Examples 11.6 and 11.3.

_ N, and ωspin are all constant, it follows that


Since, as specified in Example 11.3, θ,

dθ_
ω_ x ¼ ¼0
dt
dðN sin θÞ
ω_ y ¼ ¼ N θ_ cosθ (e)
 dt 
d ωspin + Ncosθ
ω_ z ¼ ¼ N θ_ sin θ
dt
The angular velocity Ω of the comoving xyz frame is that of the gimbal ring, which equals the angular velocity of the rotor
minus its spin. Therefore,

Ωx ¼ θ_ Ωy ¼ N sin θ Ωz ¼ N cosθ (f)


In Example 11.6, we found that
1 2 1 2 1
A¼B¼ mt + mr C ¼ mr 2 (g)
12 4 2
Substituting Eqs. (b) through (g) into Eq. (a), we get
   
1 2 1 2 1 2 1 2  _  1  
MG Þnet ¼ mt + mr ð0Þ^i + mt + mr N θ cosθ ^j + mr2 N θ_ sin θ k^
12 4 12 4 2
^i ^j ^
k
+  θ_   N sinθ  N cosθ
1 2 1 2 _ 1 2 1 2 1 2 
mt + mr θ mt + mr N sinθ mr ωspin + N cosθ
12 4 12 4 2
Expanding the cross product, collecting terms, and simplifying leads to
  
1 1 t2
MG Þnet ¼ ωspin + 3  2 N cosθ mr2 N sinθ^i
2 12 r
 2  (h)
1t 1 1
+ N cosθ  ω spin mr2 θ_ ^j  N θ_ sin θmr2 k
^
6 r2 2 2
11.7 KINETIC ENERGY 581

In Example 11.3, 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, θ ¼ 60°,
θ_ ¼ 4 rad=s, and ωspin ¼ 105 rad/s. For this set of numbers, Eq. (h) becomes

MG Þnet ¼ 0:3203^i  0:6698^j  0:1164k


^ ðN mÞ

11.7 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, ð ð
1 1
T¼ v2 dm ¼ v:vdm (11.75)
m2 m2

where v is the absolute velocity R_ of the element of mass dm. From Fig. 11.8, we infer that R_ ¼ R_ G + ρ.
_
Furthermore, Eq. (1.52) requires that ρ5ω_ _ which means
 ρ. Thus, v ¼ vG + ω  ρ,
v  v ¼ ðvG + ω  ρÞ  ðvG + ω  ρÞ ¼ vG 2 + 2vG  ðω  ρÞ + ðω  ρÞ  ðω  ρÞ
We can apply the vector identity introduced in Eq. (1.21), namely
A  ðB  C Þ ¼ B  ðC  A Þ (11.76)
to the last term to get
v  v ¼ vG 2 + 2vG  ðω  ρÞ + ω  ½ρ  ðω  ρÞ
Therefore, Eq. (11.75) becomes
ð  ð  ð
1 2 1
T¼ vG dm + vG  ω  ρdm + ω  ρ  ðω  ρÞdm
m2 m 2 m
Ð
Since the position vector ρ is measured from the center of mass, m ρdm ¼ 0. Recall that, according to
Eq. (11.34),
ð
ρ  ðω  ρÞdm ¼ HG
m

It follows that the kinetic energy may be written as


1 1
T ¼ mvG 2 + ω  HG (11.77)
2 2
The second term is the rotational kinetic energy TR,
1
TR ¼ ω  HG (11.78)
2
If the body is rotating about a point P that is at rest in inertial space, we have from Eq. (11.2) and
Fig. 11.8 that
vG ¼ vP + ω  rG=P ¼ 0 + ω  rG=P ¼ ω  rG=P

It follows that
   
vG 2 ¼ vG  vG ¼ ω  rG=P  ω  rG=P
582 CHAPTER 11 RIGID BODY DYNAMICS

Making use once again of the vector identity in Eq. (11.76), we find
   
vG 2 ¼ ω  rG=P  ω  rG=P ¼ ω  rG=P  vG

Substituting this into Eq. (11.77) yields


1
T ¼ ω  HG + rG=P  mvG
2
Eq. (11.21) shows that this can be written as
1
T ¼ ω  HP (11.79)
2
In this case, of course, all the kinetic energy is rotational.
In terms of the components of ω and H, whether it is HP or HG, the rotational kinetic energy ex-
pression becomes, with the aid of Eq. (11.39),
2 38 9
   
Ix Ixy Ixz > < ωx >
=
1 1 6 7
TR ¼ ωx Hx + ωy Hy + ωy Hz ¼ ωx ωy ωz 4 Ixy Iy Iyz 5 ωy
2 2 >
: ; >
Ixz Iyz Iz ωz

Expanding, we obtain
1 1 1
TR ¼ Ix ωx 2 + Iy ωy 2 + Iz ω2 z + Ixy ωx ωy + Ixz ωx ωz + Iyz ωy ωz (11.80)
2 2 2
Obviously, if the xyz axes are principal axes of inertia, then Eq. (11.80) simplifies considerably,

1 1 1
TR ¼ Aωx 2 + Bωy 2 + Cωz 2 (11.81)
2 2 2

EXAMPLE 11.14
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
2 3
2000 1000 2500
6 7 
½I ¼ 4 1500 3000 1500 5 kg  m2
2500 1500 4000

If at a given instant the components of angular velocity in this frame of reference are

ω ¼ 1i  0:9^j + 1:5k
^ ðrad=sÞ

calculate the total kinetic energy of the satellite.


Solution
The speed of the satellite in its circular orbit is
rffiffiffiffiffiffiffiffirffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
μ 398,600
v¼ ¼ ¼ 7:7258km=s
r 6378 + 300
11.8 THE SPINNING TOP 583

The angular momentum of the satellite is


2 38 9 8 9
2000 1000 2500 < 1 = < 6650 = 
fHG g ¼ ½IG fωg ¼ 4 1500 3000 1500 5 0:9 ¼ 5950 kg  m2 =s
: ; : ;
2500 1500 4000 1:5 9850
Therefore, the total kinetic energy is
13, 390J
44:766ð109 ÞJ
zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl
8 9ffl{
zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{ < 6650 =
1 1 1  
3 2 1
T ¼ mvG + ω  HG ¼ ð1500Þ 7:7258  10
2
+ b 1 0:9 1:5 c 5950
2 2 2 2 : ;
9850

T ¼ 44:766GJ
Obviously, the kinetic energy is dominated by that due to the orbital motion.

11.8 THE SPINNING TOP


Let us analyze the motion of the simple axisymmetric top in Fig. 11.19. It is constrained to rotate about
point O, which is fixed in space.
The moving xyz 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 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 rel-
ative to the xyz frame is diagonal, with Ix ¼ Iy ¼ A and Iz ¼ C. From Eqs. (11.68) and (11.70), we have
^i ^j ^
k
MO Þnet ¼ Aω_ x^i + Aω_ y^j + Cω_ z k
^ + Ωx Ωy Ωz (11.82)
Aωx Aωy Cωz

FIG. 11.19
Simple top rotating about the fixed point O.
584 CHAPTER 11 RIGID BODY DYNAMICS

The angular velocity ω of the top is the vector sum of the spin rate ωs and the rates of precession ωp and
nutation ωn, where
ωp ¼ ϕ_ ωn ¼ θ_ (11.83)
Thus,
ω ¼ ωn^i + ωp K
^ + ωs k
^

From the geometry, we see that


^ ¼ sinθ^j + cos θk
K ^ (11.84)
Therefore, relative to the comoving system,
 
ω ¼ ωn^i + ωp sin θ^j + ωs + ωp cos θ k^ (11.85)

From Eq. (11.85), we see that


ωx ¼ ωn ωy ¼ ωp sin θ ωz ¼ ωs + ωp cosθ (11.86)
Computing the time rates of these three expressions yields the components of angular acceleration
relative to the xyz frame, given by
ω_ x ¼ ω_ n ω_ y ¼ ω_ p sin θ + ωp ωn cos θ ω_ z ¼ ω_ s + ω_ p cos θ  ωp ωn sin θ (11.87)
^ + ωn^i, so that, using Eq. (11.84),
The angular velocity Ω of the xyz system is Ω ¼ ωp K
Ω ¼ ωn^i + ωp sinθ^j + ωp cos θk
^ (11.88)
From Eq. (11.88), we obtain
Ωx ¼ ωn Ωy ¼ ωp sinθ Ωz ¼ ωp cos θ (11.89)
In Fig. 11.19, the moment about O is that of the weight vector acting through the center of mass G:
     
^  mgK
MO Þnet ¼ d k ^  sinθ^j + cos θk
^ ¼ mgdk ^

or
MO Þnet ¼ mgd sin θ^i (11.90)
Substituting Eqs. (11.86) through (11.90) into Eq. (11.82), we get
   
mgd sin θ^i ¼ Aω_ n^i + A ω_ p sinθ + ω_ p ω_ n cosθ ^j + C ω_ s + ω_ p cos θ  ω_ p ω_ n sinθ k^
^i ^j ^
k
(11.91)
+ ωn ωp sin θ ωp cosθ
 
Aωn Aωp sinθ C ωs + ωp cosθ

Let us consider the special case in which θ is constant (i.e., there is no nutation), so that ωn ¼ ω_ n ¼ 0.
Then, Eq. (11.91) reduces to
^i ^j ^
k
 
mgd sinθ^i ¼ Aω_ p sinθ^j + C ω_ s + ω_ p cos θ k^ + 0 ωp sinθ ω
 p cos θ  (11.92)
0 Aωp sinθ C ωs + ωp cos θ
11.8 THE SPINNING TOP 585

Expanding the determinant yields


 
mgd sinθ^i ¼ Aω_ p sinθ^j + C ω_ s + ω_ p cosθ k^ + Cωp ωs sin θ + ðC  AÞωp 2 cos θ sin θ ^i

Equating the coefficients of ^i, ^j, and k


^ on each side of this equation and assuming that 0° < θ < 180°
leads to
mgd ¼ Cωp ωs + ðC  AÞωp 2 cos θ (11.93a)
Aω_ p ¼ 0 (11.93b)
 
C ω_ s + ω_ p cos θ ¼ 0 (11.93c)

Eq. (11.93b) implies ω_ p ¼ 0, and from Eq. (11.93c) it follows that ω_ s ¼ 0. Therefore, the rates of spin
and precession are both constant. From Eq. (11.93a), we find
ðA  CÞcos θωp 2  Cωs ωp + mgd ¼ 0 (11.94)

If the spin rate is zero, Eq. (11.94) yields


sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 mgd
ωp ωs ¼0 ¼  if ðC  AÞ cosθ > 0 (11.95)
ðC  AÞ cos θ

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, ωp 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 θ ¼ 0, Eq. (11.94) yields a steady precession rate that is in-
versely proportional to the spin rate,
mgd
ωp ¼ if ðA  CÞ cosθ ¼ 0 (11.96)
Cωs

If A ¼ C, this precession apparently occurs irrespective of tilt angle θ. If A 6¼ C, this rate of precession
occurs at θ ¼ 90° (i.e., the spin axis is perpendicular to the precession axis).
In general, Eq. (11.94) is a quadratic equation in ωp, so we can use the quadratic formula to find
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi!
C 4mgd ðA  CÞcos θ
ωp ¼ ωs  ωs 2  (11.97)
2ðA  CÞ cos θ C2

_
Thus, for a given spin rate and tilt angle θ (θ 6¼ 90°), there are two rates of precession ϕ.
Observe that if (A  C) cos θ > 0, then ωp is imaginary when ωs < 4mgd ðA  CÞcos θ=C2 . There-
2

fore, the minimum spin rate required for steady precession at a constant inclination θ is
2 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ωs Þmin ¼ mgd ðA  CÞcos θ if ðA  CÞ cos θ > 0 (11.98)
C
If (A  C) cos θ < 0, the radical in Eq. (11.97) is real for all ωs. In this case, as ωs ! 0, ωp approaches
the value given above in Eq. (11.95).
586 CHAPTER 11 RIGID BODY DYNAMICS

EXAMPLE 11.15
Calculate the precession rate ωp for a toy top like that in Fig. 11.19 if m ¼ 0.5 kg, A(¼Ix ¼ Iy) ¼ 12(104)kg  m2, C (¼Iz) ¼
4.5(104) kg m2, and d ¼ 0.05 m.
Solution
For an inclination of, say, 60°, (A  C) cos θ > 0, so that Eq. (11.98) requires ωs)min ¼ 407.01 rpm. Let us choose the spin
rate to be ωs ¼ 1000 rpm ¼ 104.7 rad/s. Then, from Eq. (11.97), the steady precession rate as a function of the inclination θ
is given by either one of the following formulas:
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 + 1  0:3312cosθ 1  1  0:3312cosθ
ωp ¼ 31:42 and ωp ¼ 31:42 (a)
cosθ cosθ
These are plotted in Fig. 11.20. For θ ¼ 60°, the high-energy precession rate is 1148.1 rpm, which exceeds the spin rate,
whereas the low-energy precession rate is a leisurely 51.93 rpm.

FIG. 11.20
(a) High-energy precession rate (unlikely to be observed). (b) Low-energy precession rate (the one almost
always seen).

Fig. 11.21 shows an axisymmetric rotor mounted so that its spin axis (z) remains perpendicular to
the precession axis (y). In that case, Eq. (11.85) with θ ¼ 90° yields
ω ¼ ωp^j + ωs k
^ (11.99)
Likewise, from Eq. (11.88), the angular velocity of the comoving xyz system is Ω ¼ ωp^j. If we assume
that the spin rate and precession rate are constant (dωp/dt ¼ dωs/dt ¼ 0), then Eq. (11.68), written for
the center of mass G, becomes
   
MG Þnet ¼ Ω  H ¼ ωp^j  Aωp^j + Cωs k
^ (11.100)

where A and C are the moments of inertia of the rotor about the x and z axes, respectively. Setting
^ ¼ Hs , the spin angular momentum, and ωp^j ¼ ωp , we obtain
Cωs k
MG Þnet ¼ ωp  Hs ðHs ¼ Cωs kÞ (11.101)
Since the center of mass is the reference point, there is no restriction on the motion G for which
Eq. (11.101) is valid. Observe that the net gyroscopic moment MG)net exerted on the rotor by its sup-
ports is perpendicular to the plane of the spin and the precession vectors. If a spinning rotor is forced to
11.8 THE SPINNING TOP 587

FIG. 11.21
A spinning rotor on a rotating platform.

precess, the gyroscopic moment MG)net 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 11.16
A uniform cylinder of radius r, length L, and mass m spins at a constant angular velocity ωs. It rests on simple supports
(which cannot exert couples), mounted on a platform that rotates at an angular velocity of ωp. 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 Fig. 11.22. Noting that the moment of inertia of a uniform cylinder about its axis of
rotational symmetry is mr2/2, Eq. (11.101) yields
  mr2 
RL^i ¼ ωp^j  ^ ¼ 1 mr2 ωρ ωs^i
ωs k
2 2
so that
mr2 ωρ ωs

2L

FIG. 11.22
Illustration of the gyroscopic effect.
588 CHAPTER 11 RIGID BODY DYNAMICS

11.9 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: Euler angles and yaw, pitch, and roll angles.
We will discuss each of them in turn. The reader is urged to review Section 4.5 on orthogonal coor-
dinate transformations and, in particular, the discussion of Euler angle sequences.
The three Euler angles ϕ, θ, and ψ shown in Fig. 11.23 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 ϕ. This takes X into x1 and Y into y1. The second rotation is around the
x2 (¼ x1) axis through the nutation angle θ. 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 ψ, 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 (Eq. 4.37):
½QXx ¼ ½R3 ðψ Þ½R1 ðθÞ½R3 ðϕÞ (11.102)

From Eqs. (4.32) and (4.34), we have


2 3 2 3 2 3
cos ψ sin ψ 0 1 0 0 cos ϕ sin ϕ 0
½R3 ðψ Þ ¼ 4 sin ψ cos ψ 0 5 ½R1 ðθÞ ¼ 4 0 cos θ sinθ 5 ½R3 ðϕÞ ¼ 4 sin ϕ cos ϕ 0 5 (11.103)
0 0 1 0 sin θ cos θ 0 0 1

According to Eq. (4.38), the direction cosine matrix is


2 3
 sinϕcos θ sinψ + cosϕ cos ψ cos ϕ cosθ sinψ + sinϕ cosψ sin θ sin ψ
½QXx ¼ 4 sin ϕcos θ cosγ  cos ϕsin ψ cos ϕcos θ cos ψ  sin ϕsin ψ sin θ cos ψ 5 (11.104)
sin ψ sin θ  cos ϕsin θ cosθ

FIG. 11.23
Classical Euler angle sequence (see also fig. 4.14).
11.9 EULER ANGLES 589

Since this is an orthogonal matrix, the inverse transformation from xyz to XYZ is [Q]xX ¼ [Q]TXx,
2 3
 sinϕ cos θ sinψ + cos ϕ cosψ  sinϕ cosθ cos γ  cos ϕ sinψ sinψ sinθ
½QxX ¼ 4 cos ϕcos θ sinψ + sinϕ cosψ cosϕ cos θ cos ψ  sinϕ sinψ cos ϕ sinθ 5 (11.105)
sinθ sinψ sin θ cosψ cos θ

Algorithm 4.3 is used to find the three Euler angles θ, ϕ, and ψ from a given direction cosine matrix
[Q]Xx.

EXAMPLE 11.17
The direction cosine matrix of an orthogonal transformation from XYZ to xyz is
2 3
0:32175 0:89930 0:29620
½Q ¼ 4 0:57791 0:061275 0:81380 5
0:75000 0:43301 0:5000
Use Algorithm 4.3 to find the Euler angles ϕ, θ, and ψ for this transformation.
Solution
Step 1 (precession angle, ϕ):
   
Q31 0:75000
ϕ ¼ tan 1 ¼ tan 1 ð0  ϕ < 360°Þ
Q32 0:43301
Since the numerator is negative and the denominator is positive, the angle ϕ lies in the fourth quadrant:

ϕ ¼ tan 1 ð1:7320Þ ¼ 300°


Step 2 (nutation angle, θ):

θ ¼ cos 1 Q33 ¼ cos 1 ð0:5000Þ ð0  θ  180°Þ

θ ¼ 120°
Step 3 (spin angle, ψ):
 
Q13 0:29620
ψ ¼ tan 1 ¼ tan 1 ð0  ψ < 360°Þ
Q23 0:81380
Since both the numerator and denominator are negative, the angle ψ lies in the third quadrant:

ψ ¼ tan 1 ð0:36397Þ ¼ 200°

The time rates of change of the Euler angles ϕ, θ, and ψ are, respectively, the precession rate ωp, the
nutation rate ωn, and the spin ωs. That is,
ωp ¼ ϕ_ ωn ¼ θ_ ωs ¼ ψ_ (11.106)

The absolute angular velocity ω of a rigid body can be resolved into components ωx, ωy, and ωz along
the body-fixed xyz axes, so that

ω ¼ ωx^i + ωy^j + ωz k
^ (11.107)
^ nutation is mea-
Fig. 11.23 shows that precession is measured around the inertial Z axis (unit vector K),
^
sured around the intermediate x1 axis (node line) with unit vector i1 , and spin is measured around the
590 CHAPTER 11 RIGID BODY DYNAMICS

^ Therefore, the absolute angular velocity can alternatively be written in


body-fixed z axis (unit vector k).
terms of the nonorthogonal Euler angle rates as
^ + ωn^i1 + ωs k
ω ¼ ωp K ^ (11.108)

To find the relationship between the body rates ωx, ωy, and ωz and the Euler angle rates ωp, ωn, and ωs,
we must express K ^ and ^i1 in terms of the unit vectors ^i^jk
^ of the body-fixed frame. To accomplish that,
we proceed as follows.
The first rotation [R3(ϕ)] in Eq. (11.102) rotates the unit vectors ^i^jk ^ of the inertial frame into the unit
vectors ^i1^j1 k
^1 of the intermediate x1y1z1 axes in Fig. 11.23. Hence ^i1^j1 k ^1 are rotated into ^i^jk
^ by the
inverse transformation given by
8 9 8 9 2 38 9
> ^ ^ cos ϕ sin ϕ 0 > ^
<I > = >
< i1 >
= < i1 >
=
^J ¼ ½R3 ðϕÞT ^j1 ¼ 6 7
4 sin ϕ cos ϕ 0 5 ^j1 (11.109)
>
:^> ; >
:^ > ; >
:^ ; >
K k1 0 0 1 k1
The second rotation [R1(θ)] rotates ^i1^j1 k
^1 into the unit vectors ^i2^j2 k
^2 of the second intermediate frame
x2y2z2 in Fig. 11.23. The inverse transformation rotates ^i2^j2 k^2 back into ^i1^j1 k ^1 :
8 9 8 9 2 38 9
> ^ ^ ^
< i1 >
= < i2 >
> = 1 0 0 < i2 >
> =
6 7
j1 ¼ ½R1 ðθÞ ^j2 ¼ 4 0 cos θ sin θ 5 ^j2
T
(11.110)
:^ >
> ; :^ >
> ; :^ >
> ;
k1 k2 0 sin θ cos θ k2
Finally, the third rotation [R3(ψ)] rotates ^i2^j2 k
^2 into ^i^jk,
^ the target unit vectors of the body-fixed xyz
frame. ^i2^j2 k
^2 are obtained from ^i^jk
^ by the reverse rotation,
8 9 8 9 2 38 9
> ^ ^ cos ψ sin ψ 0 > ^
< i2 >
= >
<i> = <i> =
6 7
^j2 ¼ ½R3 ðψ Þ ^j ¼ 4 sin ψ cos ψ 0 5 ^j
T
(11.111)
>
:^ > ; >
:^> ; >
:^; >
k2 k 0 0 1 k
From Eqs. (11.109) through (11.111), we observe that
11:109 11:110 11:111  
^ z}|{
K ¼ k ^ 1 z}|{ ^2 z}|{
¼ sinθ^j2 + cos θk ¼ sinθ sinψ^i + cosψ^j + cos θk
^

or
^ ¼ sinθ sinψ^i + sinθ cos ψ^j + cos θk
K ^ (11.112)
Similarly, Eqs. (11.110) and (11.111) imply that
^i1 ¼ ^i2 ¼ cos ψ^i  sinψ^j (11.113)
Substituting Eqs. (11.112) and (11.113) into Eq. (11.108) yields
   
ω ¼ ωp sin θ sinψ^i + sinθ cosψ^j + cosθk
^ + ωn cos ψ^i  sinψ^j + ωs k
^

or
     
ω ¼ ωp sinθ sin ψ + ωn cos ψ ^i + ωp sin θ cosψ  ωn sin ψ ^j + ωs + ωp cosθ k^ (11.114)
11.9 EULER ANGLES 591

Comparing the coefficients of ^i^jk


^ in this equation with those in Eqs. (11.107), we see that

ωx ¼ ωp sinθ sin ψ + ωn cos ψ


ωy ¼ ωp sinθ cos ψ  ωn sinψ (11.115a)
ωz ¼ ωs + ωp cos θ

or
8 9 2 38 9
< ωx >
> = sin θ sinψ cos ψ 0 > < ωp >
=
6 7
ωy ¼ 4 sin θ sinψ sin ψ 0 5 ωn (11.115b)
: >
> ; : >
> ;
ωz cosθ 1 ωs

(Note that the precession angle ϕ does not appear.) We solve these three equations to obtain the Euler
rates in terms of ωx, ωy, and ωz:
8 9 2 38 9
< ωp >
> = sin ψ= sinθ cos ψ= sinθ 0 >< ωx >
=
6 7
ωn ¼ 4 cos ψ  sinψ 0 5 ωy (11.116a)
>
: > ; >
: > ;
ωs sin ψ= tanθ cos ψ= tanθ 1 ωz

or
1  
ωp ¼ ϕ_ ¼ ωx sinψ + ωy cos ψ
sinθ
ωn ¼ θ_ ¼ ωx cos ψ  ωy sinψ (11.116b)
1  
ωs ¼ ψ_ ¼  ωx sinψ + ωy cos ψ + ωz
tanθ
Observe that if ωx, ωy, and ωz are given functions of time, found by solving the Euler equations of
motion (Eq. 11.72), then Eq. (11.116b) are three coupled differential equations that may be solved
to obtain the three time-dependent Euler angles, namely
ϕ ¼ ϕðtÞ θ ¼ θðtÞ ψ ¼ ψ ðtÞ

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 Eq. (11.116) “blow up” when θ ¼ 0 (i.e., when the xy plane is
parallel to the XY plane).

EXAMPLE 11.18
At a given instant, the unit vectors of a body frame are
^i ¼ 0:40825^I  0:40825^J + 0:81649K
^
^j ¼ 0:10102^I  0:90914^J  0:40405K
^ (a)
^ ¼ 0:90726^I + 0:082479^J  0:41240K
k ^

and the angular velocity is

ω ¼ 3:1^I + 2:5^J + 1:7K


^ ðrad=sÞ (b)
Calculate ωp, ωn, and ωs (the precession, nutation, and spin rates) at this instant.
592 CHAPTER 11 RIGID BODY DYNAMICS

Solution
We will ultimately use Eq. (11.116) to find ωp, ωn, and ωs. To do so we must first obtain the Euler angles ϕ, θ, and ψ 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 given unit vectors ^i, ^j, and k, ^
respectively,
2 3
0:40825 0:40825 0:81649
6 7
½QXx ¼ 4 0:10102 0:90914 0:40405 5 (c)
0:90726 0:082479 0:41240

Therefore, the components of the angular velocity in the body frame are
2 38 9 8 9
0:40825 0:40825 < 3:1 >
0:81649 > < 0:89817 >
= > =
6 7
fωgx ¼ ½QXx fωgX ¼ 4 0:10102 0:90914 0:40405 5 2:5 ¼ 2:6466
>
: > >
; : >
;
0:90726 0:082479 0:41240 1:7 3:3074

or

ωx ¼ 0:89817rad=s ωy ¼ 2:6466rad=s ωz ¼ 3:3074rad=s (d)

To obtain the Euler angles ϕ, θ, and ψ from the direction cosine matrix in Eq. (c), we use Algorithm 4.3, as was il-
lustrated in Example 11.17. 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 = [ .40825 –.40825 .81649
–.10102 –.90914 –.40405
.90726 .082479 –.41240];
[phi theta psi] = dcm_to_euler(Q)
produces the following output:
phi =
95.1945
theta =
114.3557
psi =
116.3291
Substituting θ ¼ 114.36° and ψ ¼ 116.33° together with the angular velocities of Eq. (d) into Eqs. (11.116a) and
(11.116b) yields
1
ωp ¼ ½0:89817  sin 116:33∘ + ð2:6466Þ  cos116:33∘ 
sin114:36∘
¼ 0:40492rad=s

ωn ¼ 0:89817  cos116:33∘  ð2:6466Þ  sin 116:33∘

¼ 2:7704rad=s

1
ωs ¼  ½0:89817  sin 116:33∘ + ð2:6466Þ  cos116:33∘  + ð3:3074Þ
tan114:36∘
¼ 3:1404rad=s
11.9 EULER ANGLES 593

EXAMPLE 11.19
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:

ϕ ¼ 2te0:05t
θ ¼ 0:02 + 0:3sin 0:25t (b)
ψ ¼ 0:6t
At t ¼ 10 s, find
(a) the net moment about G and
(b) the components αX, αY, and αZ of the absolute angular acceleration in the inertial frame.
Solution
(a) We must use Euler equations (Eq. 11.72) to calculate the net moment, which means we must first obtain
ωx , ωy , ωz , ω_ x , ω_ y , and ω_ z . Since we are given the Euler angles as function of time, we can compute their time
derivatives and then use Eq. (11.115) to find the body frame angular velocity components and their derivatives. Starting
with the first of Eqs. (b), we get
dϕ d  0:05t 
ωp ¼ ¼ 2te ¼ 2e0:05t  0:1te0:05t
dt dt
dωp d  0:05t 
ω_ p ¼ ¼ 2e  0:1e0:05t ¼ 0:2e0:05t + 0:005te0:05t
dt dt
Proceeding to the remaining two Euler angles leads to
dθ d
ωn ¼ ¼ ð0:02 + 0:3 sin0:25tÞ ¼ 0:075cos0:25t
dt dt
dωn d
ω_ n ¼ ¼ ð0:075 cos0:25tÞ ¼ 0:01875sin0:25t
dt dt
dψ d
ωs ¼ ¼ ð0:6tÞ ¼ 0:6
dt dt
dωs
ω_ s ¼ ¼0
dt
Evaluating all these quantities, including those in Eqs. (b), at t ¼ 10 s yields
ϕ ¼ 335:03∘ ωp ¼ 0:60653rad=s ω_ p ¼ 0:09098rad=s2

θ ¼ 11:433 ωn ¼ 0:060086rad=s ω_ ¼ 0:011221rad=s2 (c)
ψ ¼ 343:77 ωs ¼ 0:6rad=s ω_ s ¼ 0
Eq. (11.115) relates the Euler angle rates to the angular velocity components,
ωx ¼ ωp sinθ sin ψ + ωn cosψ
ωy ¼ ωp sinθ cosψ  ωn sinψ (d)
ωz ¼ ωs + ωp cosθ
Taking the time derivative of each of these equations in turn leads to the following three equations:
ω_ x ¼ ωp ωn cosθ sin ψ + ωp ωs sinθ cosψ  ωn ωs sin ψ + ω_ p sin θ sinψ + ω_ n cosψ
ω_ y ¼ ωp ωn cosθ cosψ  ωp ωs sinθsinψ  ωn ωs cosψ + ω_ p sin θ cosψ  ω_ n sinψ (e)
ω_ z ¼ ωp ωn sin θ + ω_ p cosθ + ω_ s
594 CHAPTER 11 RIGID BODY DYNAMICS

Substituting the data in Eqs. (c) into Eqs. (d) and (e) yields
ωx ¼ 0:091286rad=s ωy ¼ 0:098649rad=s ωz ¼ 1:1945rad=s
  (f)
ω_ x ¼ 0:063435rad=s2 ω_ y ¼ 2:2346 105 rad=s2 ω_ z ¼ 0:08195rad=s2
With Eqs. (a) and (f) we have everything we need for the Euler equations, namely,
Mx Þnet ¼ Aω_ x + ðC  BÞωy ωz

My net ¼ Bω_ y + ðA  CÞωz ωx
Mz Þnet ¼ Cω_ z + ðB  AÞωx ωy
from which we find
Mx Þnet ¼ 181:27 N m

My net
¼ 218:12 N m
Mz Þnet ¼ 254:86 N m
(b) Since the comoving xyz frame is a body frame, rigidly attached to the solid, we know from Eq. (11.74) that
8 9 8 9
< aX = < ω_ x =
aY ¼ ½QxX ω_ y (g)
: ; : ;
aZ ω_ z
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 Eqs. (f), onto the axes of the inertial frame. The required
orthogonal transformation matrix is given in Eq. (11.105),
2 3
sin ϕcosθ sin ψ + cosϕcosψ sinϕ cosθ cosγ  cosϕsin ψ sinϕ sinθ
6 7
½QxX ¼ 4 cosϕcosθ sinψ + sinϕcosψ cosϕcosθ cosψ  sinϕ sinψ cosϕsinθ 5
sin θ sin ψ sinθ cosψ cosθ
Upon substituting the numerical values of the Euler angles from Eqs. (c), this becomes
2 3
0:75484 0:65055 0:083668
6 7
½QxX ¼ 4 0:65356 0:73523 0:17970 5
0:055386 0:19033 0:98016
Substituting this and the relative angular velocity rates from Eqs. (f) into Eq. (g) yields
8 9 2 38 9
< αX >
> = 0:75484 0:65055 0:083668 > < 0:063435  
>
=
6 7
αY ¼ 4 0:65356 0:73523 0:17970 5 2:2345 105
>
: ; > >
: >
;
αZ 0:055386 0:19033 0:98016 0:08195
8 9
>
< 0:054755 > =
¼ 0:026716 ðrad=s2 Þ
>
: >
;
0:083833

EXAMPLE 11.20
Fig. 11.24 shows a rotating platform on which is mounted a rectangular parallelepiped shaft (with dimensions b, h, and ‘) spin-
ning about the inclined axis DE. If the mass of the shaft is m, and the angular velocities ωp and ωs are constant, calculate the
bearing forces at D and E as a function of ϕ and ψ. 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 ^I  J^ ¼ K
^ . 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 parallelepiped.
11.9 EULER ANGLES 595

FIG. 11.24
Spinning block mounted on rotating platform.

FIG. 11.25
Free body diagram of the block in Fig. 11.24.

The three Euler angles ϕ, θ, and ψ are shown in Fig. 11.24. Since θ is constant, the nutation rate is zero (ωn ¼ 0). Thus,
Eq. (11.115) reduce to

ωx ¼ ωp sin θ sin ψ ωy ¼ ωp sinθ cosψ ωz ¼ ωp cos θ + ωs (a)


Since ωp, ωs, and θ are constant, it follows (recalling Eq. 11.106) that

ω_ x ¼ ωp ωs sinθ cosψ ω_ y ¼ ωp ωs sinθ sinψ ω_ z ¼ 0 (b)


The principal moments of inertia of the parallelepiped are (Fig. 11.10C)

1  2 2 1   1  
A ¼ Ix ¼ m h +‘ B ¼ Iy ¼ m b2 + l2 C ¼ Iz ¼ m b2 + h2 (c)
12 12 12
Fig. 11.25 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.
From Newton’s laws of motion, we have Fnet ¼ maG. But G is fixed in inertial space, so aG ¼ 0. Thus,
   
Dx^i + Dy^j + Dz k
^ + Ex^i + Ey^j ¼ 0

It follows that

Ex ¼ Dx Ey ¼ Dy Dz ¼ 0 (d)


596 CHAPTER 11 RIGID BODY DYNAMICS

Summing moments about G we get


‘^  ^   ‘   
MG Þnet ¼ k  Ex i + Ey^j +  k ^  Dx^i + Dy^j + TD k
^
2 2
   
‘ ‘ ‘ ‘
¼ Dy  Ey ^i + Dx + Ex ^j + TD k ^
2 2 2 2
¼ Dy ‘^i  Dx ‘^j + TD k
^

where we made use of Eq. (d)2. Thus,



Mx Þnet ¼ Dy ‘ My net
¼ Dx ‘ Mz Þnet ¼ TD (e)
We substitute Eqs. (a) through (c) and (e) into the Euler equations (Eqs. 11.72a and 11.72b):
Mx Þnet ¼ Aω_ x + ðC  BÞωy ωz
My net ¼ Bω_ y + ðA  CÞωx ωz (f)
Mz Þnet ¼ Cω_ z + ðB  AÞωx ωy
After making the substitutions and simplifying, the first Euler equation, Eq. (f)1, becomes
1 m  2 2
Dy ¼ h  l ωp cosθ + 2h2 ωs ωp sinθ cosψ (g)
12 ‘
Likewise, from Eq. (f)2 we obtain
1 m  2 2
Dx ¼ b  l ωp cosθ + 2b2 ωs ωp sinθ sin ψ (h)
12 ‘
Finally, Eq. (f)3 yields
1  2 
TD ¼ m b  h2 ωp 2 sin 2 θ sin2ψ (i)
24
This completes the solution, since Ey ¼ Dy and Ez ¼ Dq Note that the resultant transverse bearing load V at D (and E) is
z. ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

V¼ D x 2 + Dy 2 (j)

As a numerical example, let


‘ ¼ 1m h ¼ 0:1m b ¼ 0:025m θ ¼ 30° m ¼ 10kg
and
ωp ¼ 100rpm ¼ 10:47rad=s ωs ¼ 2000rpm ¼ 209:4rad=s
For these numbers, the variation of V and TD with ψ are as shown in Fig. 11.26.

FIG. 11.26
(A) Transverse bearing load. (B) Axial torque at D.
11.10 YAW, PITCH, AND ROLL ANGLES 597

11.10 YAW, PITCH, AND ROLL ANGLES


The problem of Euler angle relations (Eq. 11.116) becoming singular when the nutation angle θ is zero
can be alleviated by using the yaw, pitch, and roll angles discussed in Section 4.5. As in the classical
Euler sequence, the yaw–pitch–roll sequence rotates the inertial XYZ axes into the triad of body-fixed
xyz axes triad by means of a series of three elementary rotations, as illustrated in Fig. 11.27. Like the
classical Euler sequence, the first rotation is around the Z (¼ z1) axis through the yaw angle ϕ. This
takes X into x1 and Y into y1. The second rotation is around the y2 (¼ y1) axis through the pitch angle
θ. This carries x1 and z1 into x2 and z2, respectively. The third and final rotation is around the x (¼ x2)
axis through the roll angle ψ, which takes y2 into y and z2 into z.
Eq. (4.40) gives the matrix [Q]Xx of the transformation from the inertial frame into the body-fixed
frame,

½QXx ¼ ½R1 ðψ Þ½R2 ðθÞ½R3 ðϕÞ (11.117)

From Eqs. (4.32) through (4.34), the elementary rotation matrices are

FIG. 11.27
Yaw, pitch, and roll sequence (see also fig. 4.15).
598 CHAPTER 11 RIGID BODY DYNAMICS

2 3 2 3
1 0 0 cosθ 0  sinθ
6 7 6 7
½R1 ðψ Þ ¼ 4 0 cos ψ sin ψ 5 ½R2 ðθÞ ¼ 4 0 1 0 5
0 sin ψ cos ψ sinθ 0 cosθ
2 3 (11.118)
cos ϕ sin ϕ 0
6 7
½R3 ðϕÞ ¼ 4 sin ϕ cos ϕ 0 5
0 0 1

According to Eq. (4.41), the multiplication on the right of Eq. (11.117) yields the following direction
cosine matrix for the yaw, pitch, and roll sequence:
2 3
cos ϕcos θ sinϕ cos θ  sinθ
6 7
½QXx ¼ 4 cosϕ sinθ sinψ  sinϕ cos ψ sin ϕsin θ sin ψ + cos ϕcos ψ cos θ sinψ 5 (11.119)
cosϕ sinθ cos ψ + sinϕ sinψ sin ϕsinθ cos ψ  cosϕ sinψ cos θ cos ψ

The inverse matrix [Q]xX, which transforms xyz into XYZ, is just the transpose
2 3
cos ϕcos θ cosϕ sinθ sinψ  sinϕ cos ψ cos ϕ sinθ cosψ + sinϕ sinψ
6 7
½QxX ¼ 4 sinϕ cos θ sinϕ sinθ sinψ + cos ϕ cosψ sinϕ sinθ sin ψ  cos ϕ sinψ 5 (11.120)
 sinθ cos θ sinψ cosθ cos ψ

Algorithm 4.4 (dcm_to_ypr.m in Appendix D.21) is used to determine the yaw, pitch, and roll angles for
a given direction cosine matrix. The following brief MATLAB session reveals that the yaw, pitch, and
roll angles for the direction cosine matrix in Example 11.17 are ϕ ¼ 109.69°, θ ¼ 17.230°, and
ψ ¼ 238.43°.
Q = [–0.32175 0.89930 –0.29620
0.57791 –0.061275 –0.81380
–0.75000 –0.43301 –0.5000];
[yaw pitch roll] = dcm_to_ypr(Q)

yaw =
109.6861
pitch =
17.2295
roll =
238.4334

Fig. 11.27 shows that yaw ϕ is measured around the inertial Z axis (unit vector K),^ pitch θ is mea-
sured around the intermediate y1 axis (unit vector ^j1 ), and roll ψ is measured around the body-fixed x
axis (unit vector ^i). The angular velocity ω, expressed in terms of the rates of yaw, pitch, and roll, is
^ + ωpitch^j2 + ωroll^i
ω ¼ ωyaw K (11.121)

in which
ωyaw ¼ ϕ_ ωpitch ¼ θ_ ωroll ¼ ψ_ (11.122)
11.10 YAW, PITCH, AND ROLL ANGLES 599

The first rotation [R3(ϕ)] in Eq. (11.117) rotates the unit vectors ^I^JK ^ of the inertial frame into the
^ ^ ^ ^ ^ ^1 are rotated into ^I^JK
unit vectors i1 j1 k1 of the intermediate x1y1z1 axes in Fig. 11.27. Thus, i1 j1 k ^ by the
inverse transformation
8 9 2 38 9
^ cosϕ  sinϕ 0 > ^
<I>
> = < i1 >
=
6 7
^J ¼ 4 sinϕ cosϕ 0 5 ^j1 (11.123)
:^>
> ; :^ >
> ;
K 0 0 1 k1
The second rotation [R2(θ)] rotates ^i1^j1 k
^1 into the unit vectors ^i2^j2 k
^2 of the second intermediate frame
x2y2z2 in Fig. 11.27. The inverse transformation rotates i2 j2 k2 back into ^i1^j1 k
^ ^ ^ ^1 :
8 9 2 38 9
> ^ cos θ 0 sinθ > ^
< i1 >
= < i2 >
=
^j1 ¼ 6
4 0 1 0 5 j2
7 ^
(11.124)
:^ >
> ; :^ >
> ;
k1  sinθ 0 cos θ k2
Lastly, the third rotation [R1(ψ)] rotates ^i2^j2 k
^2 into ^i^jk,
^ the unit vectors of the body-fixed xyz frame.
^i2^j2 k
^2 are obtained from ^i^jk
^ by the reverse transformation,
8 9 2 38 9
^ > ^
< i2 >
> = 1 0 0 <i> =
^j2 ¼ 6 7 ^
4 0 cosψ  sinψ 5 j (11.125)
:^ >
> ; :^>
> ;
k2 0 sinψ cosψ k
From Eqs. (11.123) through (11.125), we see that
11:123 11:124 11:125  
^ z}|{
K ¼ k ^1 z}|{ ^ 2 z}|{
¼ sin θ^i2 + cosθk ¼  sinθ^i + cos θ sinψ^j + cosψ k
^

or
^ ¼  sinθ^i + cos θ sinψ^j + cos θ cos ψ k
K ^ (11.126)
From Eq. (11.125),
^j2 ¼ cos ψ^j  sinψ k
^ (11.127)
Substituting Eqs. (11.126) and (11.127) into Eq. (11.121) yields
   
ω ¼ ωyaw  sinθ^i + cosθ sinψ^j + cosθ cos ψ k
^ + ωpitch cosψ^j  sinψ k
^ + ωroll^i

or
   
ω ¼ ωyaw sinθ + ωroll ^i + ωyaw cos θ sinψ + ωpitch cosψ ^j
  (11.128)
+ ωyaw cos θ cosψ  ωpitch sin ψ k^

Comparing the coefficients of ^i^jk


^ in Eqs. (11.107) and (11.128), we conclude that the body angular
velocities are related to the yaw, pitch, and roll rates as follows:
ωx ¼ ωroll  ωyaw sinθpitch
ωy ¼ ωyaw cosθpitch sin ψ roll + ωpitch cosψ roll (11.129a)
ωz ¼ ωyaw cosθpitch cos ψ roll  ωpitch sin ψ roll
600 CHAPTER 11 RIGID BODY DYNAMICS

or 8 9 2 38 9
< ωx >
> =  sinθpitch 0 1 > < ωyaw >=
6 7
ωy ¼ 4 cos θpitch sin ψ roll cos ψ roll 0 5 ωpitch (11.129b)
>
: > ; >
: >
;
ωz cos θpitch cos ψ roll sin ψ roll 0 ωroll
wherein the subscript on each symbol helps us remember the physical rotation it describes. Note that
ϕyaw does not appear. The inverse of these equations is
8 9 2 38 9
< ωyaw >
> = 0 sinψ roll = cos θpitch cosψ roll =cos θpitch >< ωx >
=
6 7
ωpitch ¼ 4 0 cos ψ roll  sinψ roll 5 ωy (11.130a)
>
: >
; : >
> ;
ωroll 1 sin ψ roll tanθpitch cos ψ roll tan θpitch ωz
or
1  
ωyaw ¼ ωy sinψ roll + ωz cos ψ roll
cos θpitch
(11.130b)
ωpitch ¼ ωy cosψ roll  ωz sinψ roll
ωroll ¼ ωx + ωy tanθpitch sinψ roll + ωz tan θpitch cosψ roll

Note that this system becomes singular (cosθpitch ¼ 0) when the pitch angle is 90°.

11.11 QUATERNIONS
In Chapter 4, we showed that the transformation from any Cartesian coordinate frame to another having
the same origin can be accomplished by three Euler angle sequences, each being an elementary rotation
about one of the three coordinate axes. We have focused on the commonly used classical Euler angle
sequence [R3(γ)][R1(β)][R3(α)] and the yaw–pitch–roll sequence [R1(γ)][R2(β)][R3(α)].
Another of Euler’s theorems, which we used in Section 1.6, states that any two Cartesian coordinate
frames are related by a unique rotation about a single line through their common origin. This line is
called the Euler axis and the angle is referred to as the principal angle.
^ be the unit vector along the Euler axis. A vector v can be resolved into orthogonal components
Let u
v? normal to u ^ and vk parallel to u^, so that we may write
v ¼ vk + v? (11.131)
^ is given by v  u
The component of v along u ^. That is,
^Þ^
vk ¼ ðv  u u (11.132)

From Eqs. (11.131) and (11.132), we have


^ Þ^
v? ¼ v  ðv  u u (11.133)
0 ^, as illustrated in Fig. 11.28.
Let v be the vector obtained by rotating v through an angle θ around u
This rotation leaves the magnitude of v? and its component along u ^ unchanged. That is
 0 
v  ¼ kv? k (11.134)
?

v0k ¼ ðv  u
^Þ^
u (11.135)
11.11 QUATERNIONS 601

FIG. 11.28
^.
Rotation of a vector through an angle θ about an axis with unit vector u

 
v0? , having been ^, has the component v0?  cos θ along the original vector v? and the
 0 rotated about u
component v?  sin θ along the vector normal to the plane of u
^ and v. Let w
^ be the unit vector normal
to that plane. Then,
v?
^ ¼u
w ^ (11.136)
kv? k
Thus,
  v?   ^  v?
u
v0? ¼ v0?  cos θ + v0?  sinθ
kv? k kv? k
According to Eq. (11.134), this reduces to
v0? ¼ cos θv? + sin θ u
^  v? (11.137)
Observe that  
u ^  v  vk ¼ u
^  v? ¼ u ^v

^. This, together with Eq. (11.133), means we can write Eq. (11.137) as
since vk is parallel to u
v0? ¼ cosθ½v  ðv  u
^Þ^ ^  vÞ
u + sinθðu (11.138)
0
Since v ¼ v0? + v0k , we find, upon substituting Eqs. (11.135) and (11.138) and collecting terms, that
v0 ¼ cosθv + ð1  cos θÞðu
^  vÞ^ ^  vÞ
u + sinθðu (11.139)
This is known as Rodrigues’ rotation formula, named for the same French mathematician who gave us
the Rodrigues’ formula for Legendre polynomials (Eq. 10.22). Eq. (11.139) is useful for determining
the result of rotating a vector about a line.
We can obtain the body-fixed xyz Cartesian frame from the inertial XYZ frame by a single rotation
through the principal angle θ about the Euler axis u ^ are thereby rotated into ^i^jk.
^. The unit vectors ^I^JK ^
The two sets of unit vectors are related by Eq. (11.139). Thus,
 
^i ¼ cos θ^I + ð1  cos θÞ u^  ^I u^ + sinθ^ u  ^I
 
^j ¼ cos θJ^ + ð1  cos θÞ u^^ J u ^ + sinθ^u^ J (11.140)
 
^ ¼ cos θK
k ^ + ð1  cosθÞ u ^K ^ u ^ + sinθ^uK ^
602 CHAPTER 11 RIGID BODY DYNAMICS

^ in terms of its direction cosines l, m, and n along the original XYZ axes.
Let us express the unit vector u
That is
^ ¼ l^I + m^J + nK
u ^ l2 + m2 + n2 ¼ 1 (11.141)
Substituting these into Eq. (11.140), carrying out the vector operations, and collecting the terms yields
^i ¼ l2 ð1  cos θÞ + cos θ ^I + ½lmð1  cosθÞ + n sinθ^ ^
J + ½lnð1  cosθÞ  m sinθK
^j ¼ ½lmð1  cos θÞ  sinθ^I + m2 ð1  cos θÞ + cos θ ^ ^
J + ½mnð1  cos θÞ + lsinθK (11.142)
^ ¼ ½lnð1  cosθÞ + msin θ^I + ½mnð1  cosθÞ  lsin θ^
k ^
J + n2 ð1  cosθÞ + cosθ K
Recall that the rows of the matrix [Q]Xx of the transformation from XYZ to xyz comprise the direction
cosines of the unit vectors ^i, ^j, and k,
^ respectively. That is,
2 3
l2 ð1  cosθÞ + cosθ lmð1  cosθÞ + n sinθ lnð1  cosθÞ  m sinθ
½QXx ¼ 4 lmð1  cosθÞ  n sin θ m2 ð1  cos θÞ + cos θ mnð1  cos θÞ + lsin θ 5 (11.143)
lnð1  cos θÞ + msinθ mnð1  cos θÞ  lsinθ n2 ð1  cos θÞ + cos θ
The direction cosine matrix is thus expressed in terms of the Euler axis direction cosines and the
principal angle.
Quaternions (also known as Euler symmetric parameters) were introduced in 1843 by the Irish
mathematician Sir William R. Hamilton (1805–65). They provide an alternative to the use of direction
cosine matrices for describing the orientation of a body frame in three-dimensional space. Quaternions
can be used to avoid encountering the singularities we observed for the classical Euler angle sequence
when the nutation angle θ becomes zero (Eqs. 11.116a and 11.116b) or for the yaw, pitch, and roll
sequence when the pitch angle θ approaches 90° (Eq. (11.126)).
_
As the name implies, a quaternion q comprises four numbers:

q1
q2 q
q= = (11.144)
q3 q4
q4

 
where q is called the vector part q ¼ q1^i + q2^j + q3 k
^ , and q4 is the scalar part. (It is common to see the

scalar part denoted q0 and listed first, in which case .) Regardless, a quaternion whose scalar
part is zero is called a pure quaternion.
_ _
The norm q  of the quaternion q is defined as
_ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
q  ¼ kqk2 + q2 ¼ q  q + q2 (11.145)
4 4

Obviously, the norm of a pure


_quaternion
 (q4 ¼ 0) is just the norm of its vector part. A unit quaternion is
one whose norm is unity (q  ¼ 1).
Quaternions obey the familiar vector rules of addition and scalar multiplication. That is,

p +q ap
p +q = ap =
p4 + q4 ap4
11.11 QUATERNIONS 603

Addition is both associative and commutative, so that


   
_ _ _ _ _ _ _ _ _ _
p +q +r ¼p + q +r p +q ¼q +p

We use the special symbol to denote the product or “composition” of two quaternions. The somewhat
complicated rule for quaternion multiplication involves ordinary scalar multiplication as well as the
familiar vector dot product and cross product operations,

p4q + q4 p + p q
p q= (11.146)
p4 q4 − p q

Switching the order of multiplication yields

q4 p + p4q + q p
q p=
q4 p4 − q p

We are familiar with the fact that q  p ¼ (p  q), which means that quaternion multiplication is
generally not commutative,
_ _ _ _
p q 6¼ q p

EXAMPLE 11.21
Find the product of the quaternions

p j q 0.5i + 0.5j + 0.75k


p q= = (a)
p4 1 q4 1

Solution

p4q + q4p + p × q
p q =
p4q4 – p q

0.5i + 0.5j + 0.75k j –0.5k + 0.75i


1 0.5i + 0.5j + 0.75k + 1 j + j × 0.5i + 0.5j + 0.75k
=
1 1 – j 0.5i + 0.5j + 0.75k
1 0.5
0.5 + 0.75 i + 0.5 + 1.0 j + 0.75 – 0.5 k
=
0.5
604 CHAPTER 11 RIGID BODY DYNAMICS

or

p q = 1.25i + 1.5j + 0.25k


0.5

_ _
The conjugate q * of a quaternion q is found by simply multiplying its vector part by 1, thereby
changing the signs of its vector components:

–q
q*= (11.147)
q4

_
The identity quaternion 1 has zero for its vector part and 1 for its scalar part,

0
1= (11.148)
1
_
The product of any quaternion with 1 is commutative and yields the original quaternion,

q4 0 + 1 q + q × 0 q
q 1 1 q q
q4 1 – q 0 q4

Multiplication of_a quaternion by its conjugate is also a commutative operation that yields a quaternion
proportional to 1 ,

q4 q q4 q q q 0
q q q q q 1 (11.149)
q4q4 q q q

_1
The inverse q of a quaternion is defined as
_
_1 q*
q ¼ _2 (11.150)
q 
_ _2 _1
Substituting q * ¼ q  q into Eq. (11.149) yields

_ _1 _1 _ _
q q ¼q q ¼1 (11.151)
_ _1
Clearly, for unit quaternions the inverse and the conjugate are the same, q * ¼ q , and
_    
_ _ _ _ _
q q* ¼ q* q ¼ 1 if q  ¼ 1 (11.152)
11.11 QUATERNIONS 605

Let us restrict our attention to unit quaternions, in which case

sin q /2 u
q (11.153)
cos q /2

^ is the unit vector along the Euler axis around which the inertial reference frame is rotated into
where u
the body-fixed frame, and θ is the Euler principal rotation angle. Recalling Eq. (11.141), we observe
that
q1 ¼ lsin ðθ=2Þ q2 ¼ msin ðθ=2Þ q3 ¼ n sin ðθ=2Þ q4 ¼ cos ðθ=2Þ (11.154)
_ _
The conjugate quaternion q * is found by reversing the sign of the vector part of q , so that

– sin q /2 u
q (11.155)
cos q /2

Employing these and the trigonometric identities


cos θ ¼ cos 2 ðθ=2Þ  sin 2 ðθ=2Þ sinθ ¼ 2cos ðθ=2Þ sin ðθ=2Þ (11.156)

we show in Appendix G that the direction cosine matrix [Q]Xx of the body frame in Eq. (11.143) is
_
obtained from the quaternion q by means of the following algorithm.

ALGORITHM 11.1
_
Obtain the direction cosine matrix [Q]Xx from the unit quaternion q . This procedure is implemented
in the MATLAB function dcm_from_q.m in Appendix D.49.
1. Write the quaternion as

q1
q2
q
q3
q4
_
where b q1 q2 q3 cT is the vector part, q4 is the scalar part, and q  ¼ 1.
2. Compute the direction cosine matrix of the transformation from XYZ to xyz as follows:
2 3
q21  q22  q23 + q24 2ðq1 q2 + q3 q4 Þ 2 ð q1 q3  q 2 q4 Þ
½QXx ¼ 4 2ðq2 q1  q3 q4 Þ q21 + q22  q23 + q24 2ðq2 q3 + q1 q4 Þ 5 (11.157)
2ðq3 q1 + q2 q4 Þ 2ðq3 q2  q1 q4 Þ q21  q22 + q23 + q24
606 CHAPTER 11 RIGID BODY DYNAMICS

Note that every element of the matrix [Q]Xx in Eq. (11.157) contains products of two components
_ _ _
of q . Since q and q therefore yield the same direction cosine matrix, they represent the same rotation.
We can verify by carrying out the matrix multiplication and using Eq. (11.145) that [Q]Xx in Eq. (11.157)
exhibits the required orthogonality property,
½QXx ½QXx T ¼ ½QXx T ½QXx ¼ ½1
To find the unit quaternion (q21 + q22 + q23 + q24 ¼ 1) for a given direction cosine matrix, we observe
from Eq. (11.157) that
1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
q4 ¼ 1 + Q11 + Q22 + Q33
2
(11.158)
Q23  Q32 Q31  Q13 Q12  Q21
q1 ¼ q2 ¼ q3 ¼
4q4 4q4 4q4
This procedure obviously fails for pure quaternions (q4 ¼ 0). The following algorithm (Bar-Itzhack,
2000) avoids having to deal with this situation.

ALGORITHM 11.2
Obtain the (unit) quaternion from the direction cosine matrix [Q]Xx. This procedure is implemented
as the MATLAB function q_ from_dcm.m in Appendix D.50.
1. Form the 4-by-4 symmetric matrix
2 3
Q11  Q22  Q33 Q21 + Q12 Q31 + Q13 Q23  Q32
1 6 Q21 + Q12 Q11 + Q22  Q33 Q32 + Q23 Q31  Q13 7
½K ¼ 64
7 (11.159)
3 Q 31 + Q 13 Q 32 + Q23 Q 11  Q 22 + Q 33 Q12  Q21 5
Q23  Q32 Q31  Q13 Q12  Q21 Q11 + Q22 + Q33
2. Solve the eigenvalue problem [K]{e} ¼ λ{e}nforo the largest eigenvalue λmax. The
_
corresponding eigenvector is the quaternion, q ¼ feg. Since we are interested in only the
dominant eigenvalue of [K], we can use the iterative power method (Jennings, 1977), which
converges to λmax. Thus, starting with an estimate {e0} of the eigenvector, we normalize it,
fe 0 g
f^e0 g ¼
ke 0 k
and use f^e0 g to compute an updated normalized estimate
½Kf^e0 g
f^e1 g ¼ ðk^e1 k ¼ 1Þ
k½Kf^e0 gk
We estimate the corresponding eigenvalue by using the Rayleigh quotient,

f^e1 gT ½Kf^e1 g
r1 ¼ ¼ f^e1 gT ½Kf^e1 g
k^e1 k 2

We repeat this process, using ^e1 to compute an updated normalized estimate ^e2 followed
by using the Rayleigh quotient to find r2, and so on, over and over again. After n steps
11.11 QUATERNIONS 607

we have ^en and rn. As n increases, rn converges toward λmax, the maximum eigenvalue of [K].
When j(rn  rn1)/rn1 j < ε, where ε is our chosen tolerance, we terminate the iteration and
declare that λmax ¼ rn and that ^en is the corresponding eigenvector.
Of course, instead of the power method we can take advantage of commercial software,
such as MATLAB’s eigenvalue extraction program eig.

EXAMPLE 11.22
(a) Write down the unit quaternion for a rotation about the x axis through an angle θ.
(b) Obtain the corresponding direction cosine matrix.
Solution
(a) According to Eq. (11.151),

q ¼ sin ðθ=2Þ^i q4 ¼ cos ðθ=2Þ (a)


so that

sin q / 2

q 0 (b)
0
cos q / 2

(b) Substituting q1 ¼ sin (θ/2), q2 ¼ q3 ¼ 0, and q4 ¼ cos (θ/2) into Eq. (11.152) yields
2 2 3
sin ðθ=2Þ + cos 2 ðθ=2Þ 0 0
6 7
½Q ¼ 4 0 sin 2 ðθ=2Þ + cos 2 ðθ=2Þ 2sin ðθ=2Þcos ðθ=2Þ 5 (c)
0 2sin ðθ=2Þcos ðθ=2Þ sin ðθ=2Þ + cos ðθ=2Þ
2 2

From trigonometry, we have


θ θ θ θ θ θ
sin 2 + cos 2 ¼ 1 2sin cos ¼ sin θ cos 2  sin 2 ¼ cosθ
2 2 2 2 2 2
Therefore, Eq. (c) becomes
2 3
1 0 0
½Q ¼ 4 0 cosθ sinθ 5 (d)
0 sin θ cosθ
We recognize this as the direction cosine matrix [R1(θ)] for a rotation θ around the x axis (Eq. 4.33).

EXAMPLE 11.23
For the yaw–pitch–roll sequence ϕyaw ¼ 50°, θpitch ¼ 90°, and ψ roll ¼ 120°, calculate
(a) the quaternion and
(b) the rotation angle and the axis of rotation.
608 CHAPTER 11 RIGID BODY DYNAMICS

Solution
(a) Substituting the given angles into Eq. (11.119) yields the direction cosine matrix
2 3
0 0 1
6 7
½QXx ¼ 6
4 0:93969 0:34202 0 5
7 (a)
0:34202 0:93969 0
Substituting the components of [Q]Xx into Eq. (11.159), we get
2 3
0:11401 0:31323 0:21933 0:31323
6 7
6 0:31323 0:11401 0:31323 0:44734 7
6 7
½K ¼ 6 7 (b)
6 0:21933 0:31323 0:11401 0:31323 7
4 5
0:31323 0:44734 0:31323 0:11401
The following is a MATLAB script that implements the power method described in Algorithm 11.2.
K = [-0.11401 0.31323 -0.21933 0.31323
0.31323 0.11401 -0.31323 0.44734
-0.21933 -0.31323 -0.11401 -0.31323
0.31323 0.44734 -0.31323 0.11401];
v0 = [1 1 1 1]0 ; %Initial estimate of the eigenvector.
v0 = v0/norm(v0); %Normalize it.
lamda_new = v00 *K*v0; %Rayleigh quotient (norm(v0) = 1)
% estimate of the eigenvalue.
lamda_old = 10*lamda_new; %Just to begin the iteration.
no_iterations = 0; %Count the number of iterations.
tolerance = 1.e-10;
while abs((lamda_new - lamda_old)/lamda_old) > tolerance
no_iterations = no_iterations + 1;
lamda_old = lamda_new;
v = v0;
vnew = K*v/norm(K*v);
lamda_new = vnew0 *K*vnew; %Rayleigh quotient (norm(vnew) = 1).
v0 = vnew;
end
no_iterations = no_iterations
disp(‘ ‘)
lamda_max = lamda_new
eigenvector = vnew

The output of this program to the Command Window is as follows:


no_iterations =
12
lamda_max =
1
eigenvector =
0.40558
0.57923
-0.40558
0.57923
11.11 QUATERNIONS 609

The quaternion is the eigenvector associated with λmax, so that

0.40558
q 0.57923
0.40558
0.57923

_ _
Observe that q  ¼ 1. q must be a unit quaternion.
(b) From Eq. (11.146), we find that the principal angle is

θ ¼ 2cos 1 ðq4 Þ ¼ 2cos 1 ð0:57923Þ ¼ 54:604°

and the Euler axis is

0:40558^I + 0:57923^J  0:40558K


^

u ¼ 0:4975^I + 0:71056^J  0:49754K
^
sin ð54:604°=2Þ

^
We have seen that a unit quaternion of Eq. (11.153) represents a rotation about the unit vector u
through the angle θ. Let us show that the Rodrigues’ formula (Eq. 11.139) for rotating the vector v
into the vector v0 may be written in terms of quaternions as follows:

_0 _ _ _
v ¼ q v q* (11.160)

_ _ _
where the conjugate quaternion q * is given by Eq. (11.155), and v and v 0 are the pure quaternions
having v and v0 as their vector parts,

v v v v (11.161)
0 0

Eq. (11.160) is implemented in MATLAB as the function quat_rotate.m in Appendix D.51.


_ _
Using Eq. (11.146) we first calculate the product of q and v ,

cos q / 2 v sin q / 2 u v
q v
–sin q / 2 u v

_
Multiply this quaternion on the right by q * to get

v
v q v q (11.162)
4

_
where v0 and v40 are, respectively, the vector and scalar parts of the quaternion v 0 .
610 CHAPTER 11 RIGID BODY DYNAMICS

Once again we employ Eq. (11.146) to obtain


ð_q _v Þ4 q* q*4 q v
zfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflffl{ zfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflffl{ zfflfflfflfflffl}|fflfflfflfflffl{ zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{
0
v ¼ ½ sin ðθ=2Þðu ^  vÞ ð sin ðθ=2Þ^ uÞ + cos ðθ=2Þ ½ cos ðθ=2Þv + sin ðθ=2Þðu ^  vÞ
+ ½ cos ðθ=2Þv + sin ðθ=2Þðu ^  vÞ ð sin ðθ=2Þ^ uÞ
|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl} |fflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflffl}
q v q*

¼ sin ðθ=2Þ^
2
^  vÞ + cos ðθ=2Þv + cos ðθ=2Þ sin ðθ=2Þðu
uðu 2
^  vÞ
^  vÞ  u
^  vÞ  sin 2 ðθ=2Þðu
+ cos ðθ=2Þ sin ðθ=2Þðu ^

^  vÞ  u
According to the bac–cab rule (Eq. 1.20), ðu ^ ¼vu
^ðu
^  vÞ. Substituting this into the above
equation and collecting terms we get
v0 ¼ v cos 2 ðθ=2Þ  sin 2 ðθ=2Þ + u
^ ðu
^  vÞ 2 sin 2 ðθ=2Þ + ðu
^  vÞ½2 cos ðθ=2Þ sin ðθ=2Þ

But
cos 2 ðθ=2Þ  sin 2 ðθ=2Þ ¼ cosθ 2 sin 2 ðθ=2Þ ¼ 1  cosθ 2 cos ðθ=2Þ sin ðθ=2Þ ¼ sin θ
so that finally
v0 ¼ v cosθ + u
^ ðu
^  vÞð1  cos θÞ + ðu
^  vÞ sin θ (11.163)
According to Eq. (11.146), the scalar part v4 of the quaternion product in Eq. (11.162) is
ð_q _v Þ4 q*4 q v q*
zfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflffl{ zfflfflfflfflfflfflffl}|fflfflfflfflfflfflffl{ zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{ zfflfflfflfflfflfflfflfflffl}|fflfflfflfflfflfflfflfflffl{
v04 ¼ ½ sin ðθ=2Þðu ^  vÞ ½ cos ðθ=2Þ  ½ cos ðθ=2Þv + sin ðθ=2Þðu ^  vÞ  ½ sin ðθ=2Þ^ u
^  vÞ + cos ðθ=2Þ sin ðθ=2Þðu
¼  sin ðθ=2Þ cos ðθ=2Þðu ^  vÞ  u
^  vÞ + sin 2 ðθ=2Þðu ^
¼0
_ _
Thus, the scalar part of v 0 vanishes, which means that v 0 is a pure quaternion whose vector part v0 is
_ _ _
identical to Eq. (11.139). We have therefore shown that the quaternion operation q v q * indeed
rotates the vector v around the axis of the quaternion (the Euler axis) through the angle θ. In the same
_ _ _
way we can show that the operation q * v q rotates the vector v through the angle  θ. In fact, if we
_ _ _ _ _ _
follow the operation q v q * (rotation through +θ) with the operation q * v q (rotation through
_
 θ) we end up where we started (namely, with the pure quaternion v ):
      _ _  _
_ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _
q q* v q q* ¼ q q* v q q* ¼ 1 v 1 ¼ 1 v 1 ¼v 1 ¼v

The operation in Eq. (11.160) is a vector rotation. The frame of reference remains fixed while the
vector v is rotated into the vector v0 . On the other hand, the familiar operation {v}x0 ¼ [Q]xx0 {v} is a
frame rotation (a coordinate transformation), in which the vector v remains fixed while the reference
frame is rotated. We can easily illustrate this by revisiting Example 11.22.

EXAMPLE 11.24
Consider the vector v ¼ v^j. Using the quaternion and corresponding direction cosine matrix in Example 11.22, carry out the
following operations and interpret the results geometrically:
_ _ _ _
(i) v 0 ¼ q v q *
11.11 QUATERNIONS 611

(ii) fv0 g ¼ ½Qfvg


where

sin q / 2 i 1 0 0
j
v q Q 0 cos q sin q
cos q / 2
0 sinq cosq

Solution
(i) We do the two quaternion products one after the other using Eq. (11.146). The first product is

cos q / 2 j 0 sin q / 2 i sin q / 2 i j


q v
cos q / 2 0 sin q / 2 i j

cos q / 2 j sin q / 2 k
0

Following this by the second product, we get

q v q cos q 2 j sin q 2 k sin q 2 i


0 cos q 2

0 sin q 2 i cos q 2 cos q 2 j sin q 2 k cos q 2 j sin q 2 k sin q 2 i


0 cos q 2 cos q 2 j sin q 2 k sin q 2 i

cosq sinq
cos q j sin q k
cos q 2 sin q 2 j 2 sin q 2 cos q 2 k
0
0

Finally, therefore,

v0 ¼ v cosθ^j + vsinθk
^ (a)

(ii)
2 38 9 8 9
1 0 0 <0= < 0 =
fv g ¼ 4 0 cosθ sinθ 5 v ¼ v cosθ
0
: ; : ;
0 sin θ cosθ 0 v sinθ
or

v0 ¼ v cosθ^j  v sin θk
^ (b)

These two results are illustrated in Fig. 11.29.


612 CHAPTER 11 RIGID BODY DYNAMICS

FIG. 11.29
Vector rotation vs frame rotation.

The Euler equations of motion for a rigid body (Eq. 11.72) provide the angular velocity rates ω_ x , ω_ y ,
and ω_ z as functions of time. We can integrate those equations to find the time history of the angular ve-
locities ωx, ωy, and ωz. In addition, to obtain the orientation history of the body, we need the time history of
the Euler angles ϕ, θ, and ψ. These are found by integrating Eq. (11.116),
2 3
8 9 sinψ cos ψ 8 9
> _
ϕ > 6 sinθ 0 7> ωx >
>
< > = 6 sin θ 7>< > =
6 7
θ_ ¼ 6 cos ψ sin ψ 0 7 ωy
>
> > 6 7>> >
: > ; 4 sinψ cos ψ 5: ωz ;
>
ψ_   1
tan θ tan θ

which provide the Euler angle rates (precession, nutation, and spin) in terms of the angular velocities. If
we elect to use quaternions instead of Euler angles to describe the attitude of the body, then we need a
_
formula for the rate of change of q in terms of the angular velocities.
_
To find the time derivative of a pure quaternion q , we simply differentiate Eq. (11.153) to get

d
u sin q 2 u sin q 2 u q 2 cos q 2
dt
q
d
cos q 2 q 2 sin q 2
dt

^ is constant in magnitude, but not in direction. According to Gelman (1971)


The Euler axis unit vector u
and Hughes (2004), its time derivative is
1
^_ ¼ ½u
u ^  ω  cot ðθ=2Þ^ ^  ωÞ
u  ðu (11.164)
2

where ω is the angular velocity vector. Clearly, if the instantaneous axis of rotation and the Euler axis
happen to coincide (i.e., if ω ¼ ω^ ^_ ¼ 0, because in that case u
u), then u ^  ω ¼ 0. However, in general
11.11 QUATERNIONS 613

^_ does not vanish. Substituting Eq. (11.164) into the expression for q_ , we find after expanding and
_
u
collecting terms that
1 sin q 2 u w cos q 2 w
q= (11.165)
2 q sin q 2

^ sin ðθ=2Þ ¼ q and cos(θ/2) ¼ q4. Observing furthermore that θ_ is the


From Eq. (11.153) we know that u
^), the expression for q_ becomes
_
component of the angular velocity ω along the Euler axis (θ_ ¼ ω  u

1 q w q4w
q= (11.166)
2 w q

According to the quaternion composition rule (Eq. 11.146), this may be written
_
_ 1_ _
q¼ q ω
2

where is the pure quaternion version of the angular velocity vector. Substituting

q ¼ q1^i + q2^j + q3 k
^ and ω ¼ ω1^i + ω2^j + ω3 k
^ into Eq. (11.166) and expanding the vector and scalar prod-
ucts leads to
q2w3 – q3w2 + q4w1 0 w3 –w2 w1 q1
1 q3w1 – q1w3 + q4w2 1 –w3 0 w1 w2 q2
q= =
2 q1w2 – q2w1 + q4w3 2 w2 –w1 0 w3 q3
–w1q1 – w2q2 – w3q3 –w1 –w2 –w3 0 q4

That is,
d n_o 1 n_o
q ¼ ½Ω q (11.167)
dt 2
where
0 3 2 1

3 0 1 2
(11.168)
2 1 0 3

1 2 3 0

ω1, ω2, and ω3 are the x, y, and z components of angular velocity in the body-fixed frame.
If the components of the angular velocity are constant, then the matrix [Ω] is constant and we can
readily integrate Eq. (11.167) to obtain
n o  
_ ½Ω n_ o
q ¼ exp t q0 (11.169)
2
614 CHAPTER 11 RIGID BODY DYNAMICS

n o
_
where q 0 is the value of the quaternion at time t ¼ 0. This expression may be inferred directly from
scalar calculus, in which we know that if c is a constant, then the solution of the differential equation
dx/dt ¼ cx is simply x ¼ x0ect. In linear algebra we learn that a 4-by-4 matrix [A] has four eigenvalues λi
and four corresponding eigenvectors {ei}, satisfying the equation

½Afei g ¼ λi fei g i ¼ 1,…,4

It turns out that

exp ð½AÞ ¼ ½V exp ð½ΛÞ½V1 (11.170)

where [Λ] is the 4-by-4 diagonal matrix of eigenvalues,


2 3
λ1 0 0 0
60 λ2 0 07
½Λ ¼ 6
40
7
0 λ3 05
0 0 0 λ4

and [V] is the 4-by-4 matix whose columns comprise the four distinct eigenvectors {ei},

½V ¼ ½ fe1 g fe2 g fe3 g fe4 g 

Using, for example, MATLAB’s symbolic math feature, we find that the eigenvalues and correspond-
ing eigenvectors of the matrix [Ω] in Eq. (11.168) are
8   9 8   9
>
> ωy ωi  ωx ωz =ωxy 2 >> >
>  ωx ωi + ωy ωz =ωxy 2 >>
>  
<  ω ωi + ω ω =ω 2 = > > 
< ω ωi + ω ω =ω 2 > =
x y z xy y x z xy
λ1 ¼ λ2 ¼ ωi: e1 ¼ e2 ¼
>
> 1 >
> >
> 0 >
>
>
: >
; >
: >
;
0 1
8   9 8   9
>
>  ωy ωi + ωx ωz =ωxy 2 >
> >
> ωx ωi  ωy ωz =ωxy 2 >
>
> 
<  2 =
> >
<   >
ωx ωi  ωy ωz =ωxy ωy ωi + ωx ωz =ωxy 2 =
λ3 ¼ λ4 ¼ ωi: e3 ¼ e4 ¼
>
> 1 >
> >
> 0 >
>
>
: >
; >
: >
;
0 1
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
where ωffi ¼ ωx 2 + ωy 2 + ωz 2 (the magnitude of the angular velocity vector), ωxy ¼ ωx 2 + ωy 2 , and
pffiffiffiffiffiffi
i ¼ 1. Substituting these results into Eq. (11.170) yields, again with the considerable aid of
MATLAB,
2 ωt ωz ωt ωy ωt ωx ωt 3
cos sin  sin sin
6 2 ω 2 ω 2 ω 27
  6 ωz ωt ωt ωx ωt ωy ωt 7
½Ω  6  sin cos sin sin 7
6
t ¼ 6 ωωy 2 2 ω 2 ω 27
exp ωt ωx ωt ωt ωz ωt 7 (11.171)
2 6  sin sin 7
6 ω sin 2 27
cos
4 ω 2 2 ω 5
ωx ωt ωy ωt ωz ωt ωt
 sin  sin  sin cos
ω 2 ω 2 ω 2 2
11.11 QUATERNIONS 615

We know that for Eq. (11.171) to be valid, the angular velocity components must all be constant.
The rigid body equations of motion (Eq. 11.72) show that ω_ x ¼ ω_ y ¼ ω_ z ¼ 0 if the net torque on the
body is zero and the principal moments of inertia are all the same. Whereas torque-free motion
(Chapter 12) is quite common for space vehicles, spherical symmetry (A ¼ B ¼ C) is not. Thus, we
cannot make much practical use of Eqs. (11.169) and (11.171). In general, we must instead use numer-
ical integration to obtain the angular velocities from the Euler equations and the quaternions from
Eq. (11.167).

EXAMPLE 11.25
At time t ¼ 0 the body-fixed axes and inertial angular velocity of a rigid body are those given in Example 11.18, namely
^i0 ¼ 0:40825^I  0:40825^J + 0:81649K
^
^j0 ¼ 0:10102^I  0:90914J^  0:40405K^ (a)
^0 ¼ 0:90726^I + 0:082479^J  0:41240K
k ^

and

ωX ¼ 3:1^I + 2:5^J + 1:7K


^ ðrad=sÞ (b)
If the angular velocity is constant, find the time histories of the Euler angles and the quaternion.
Solution
Because the angular velocity is constant, the motion of the body will be pure rotation about the fixed axis of rotation defined
by the angular velocity vector. Once we find the direction cosine matrix as a function of time, we can use Algorithm 4.3 to
obtain the Euler angles at each time.
Step 1:
As in Example 11.18, we find that the direction cosine matrix at time t ¼ 0 is
2 3
0:40825 0:40825 0:81649
½Q0 Xx ¼ 0:10102 0:90914 0:40405 5
4 (c)
0:90726 0:082479 0:41240
Step 2:
As in Example 11.18, use [Q0]Xx to project the angular velocity ωX onto the axes of the body-fixed frame
2 38 9
0:40825 0:40825 0:81649 < 3:1 =
fωgx ¼ ½Q0 Xx fωgX ¼ 4 0:10102 0:90914 0:40405 5 2:5
: ;
0:90726 0:082479 0:41240 1:7
so that
ωx ¼ 0:89817rad=s ωy ¼ 2:6466rad=s ωz ¼ 3:3074rad=s (d)
The magnitude of the constant angular velocity is
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ω ¼ ωx 2 + ωy 2 + ωz 2 ¼ 4:3301rad=s (e)

The period of the rotation is T ¼ 2π/ω ¼ 1.451 s.


Step 3:
Use the angular velocities in (d) to form the matrix [Ω] in Eq. (11.168),
2 3
0 3:3074 2:6466 0:89817
6 3:3074 0 0:89817 2:6466 7
½Ω  ¼ 6
4 2:6466
7 (f)
0:89817 0 3:3074 5
0:89817 2:6466 3:3074 0
616 CHAPTER 11 RIGID BODY DYNAMICS

[Ω] remains constant.


Step 4:
Use Algorithm 11.2 to obtain the quaternion at t ¼ 0 from the direction cosine matrix in (c).

–0.82610
q0 0.15412 (g)
–0.52165
0.14724
Step 5
At each time through tfinal:
_
Compute the quaternion q ðtÞ from Eqs. (11.169) and (11.171).
_
Use q ðtÞ to update the direction cosine matrix [Q(t)]Xx using Algorithm 11.1.
Use [Q(t)]Xx to calculate the Euler angles ϕ(t), θ(t), and ψ(t)by means of Algorithm 4.3.
_
Fig. 11.30 shows the time variation of the four components of q during one rotation of the body. The variation of the
three Euler angles is shown in Fig. 11.31. Observe that their values at t ¼ 0 agree with those found in Example 11.18.
Fig. 11.32 shows the initial orientation of the orthonormal body-fixed xyz axes given in Eq. (a). The dotted lines trace
out the subsequent motion of their end points as they rotate at 4.33 rad/s about the fixed angular velocity vector ω. Finally,
^ and its subsequent motion during rotation of the body. The unit
Fig. 11.33 shows the initial position of the Euler axis u
_
^ is obtained from the unit quaternion q ðtÞ at any instant by means of Eq. (11.146),
vector u
qðtÞ
^ðtÞ ¼
u
sin f cos 1 ½q4 ðtÞg
_
^ and u
where q(t) is the vector part of q ðtÞ. Fig. 11.33 amply illustrates the fact that the unit vectors ω ^ are not the same.

FIG. 11.30
_
History of the components of q for one rotation of the body.
11.11 QUATERNIONS 617

FIG. 11.31
History of the three Euler angles for one rotation of the body.

FIG. 11.32
The motion of three orthogonal lines during rotation of the body.
618 CHAPTER 11 RIGID BODY DYNAMICS

FIG. 11.33
Motion of the Euler axis during one rotation of the body.

EXAMPLE 11.26
Solve the spinning top problem of Example 11.15 numerically, using quaternions. Use the low-energy precession rate,
ωp ¼ 51.93 rpm.
Solution
We will use Eq. (11.72) (the Euler equations) to compute the body frame angular velocity derivatives:
dωx Mx C  B
¼  ωy ωz
dt A A
dωy My A  C
¼  ωz ωx (a)
dt B B
dωz Mz B  A
¼  ωx ωy
dt C C
These require that the moving xyz axes are all rigidly attached to the top. In Example 11.15 only the x axis was fixed to the
top along its spin axis; the y and z axes did not rotate with the top.
11.11 QUATERNIONS 619

The moments in Eq. (a) must be expressed in components along the body-fixed axes. From Fig. 11.19, the moment of
the weight vector about O is
   
^  mgK
M ¼ dk ^ ¼ mgd k^K
^ (b)

where, recalling Eq. (4.18)3

^ ¼ Q31^I + Q32 ^J + Q33 K


k ^ (c)

The Qs are the time-dependent components of the direction cosine matrix [Q]Xx in Eq. (11.157). Carrying out the cross
product in Eq. (b) yields the components of M along the XYZ axes of the fixed space frame,

8 9
< mgdQ32 =
fMgX ¼ mgdQ31
: ;
0

To obtain the components of M in the body-fixed frame, we perform the transformation


2 38 9 8 9
Q11 Q12 Q13 < mgdQ32 = < Q12 Q31  Q32 Q11 =
6 7
fMgx ¼ ½QXx fMgX ¼ 4 Q21 Q22 Q23 5 mgdQ31 ¼ mgd Q22 Q31  Q32 Q21 (d)
: ; : ;
Q31 Q32 Q33 0 0

It can be shown that (Problem 11.27)

Q12 Q31  Q32 Q11 ¼ Q23 Q22 Q31  Q32 Q21 ¼ Q13

Therefore, at any instant the moments in Eq. (a) are

Mx ¼ mgdQ23 My ¼ mgdQ13 Mz ¼ 0 (e)

The MATLAB implementation of the following procedure is listed in Appendix D.51.


Step 1:
Specify the initial orientation of the xyz axes of the body frame, thereby defining the initial value of the direction cosine
matrix [Q]Xx.
According to Fig. 11.19, the body z axis is the top’s spin axis, and we shall assume here that it initially lies in the global
YZ plane, tilted 60° away from the Z axis, as it is in Example 11.15. Let the body x axis be initially aligned with the global X
axis. The body y axis is then found from the cross product ^j ¼ k ^ ^i. Thus,

^ ¼ sin 60°^J + cos60°K


k ^
^i ¼ ^I
^j ¼ k
^ ^i ¼ cos60°^J + sin60°K
^

It follows that the direction cosine matrix relating XYZ to xyz at the start of the simulation is
2 3 2 3
1 0 0 1 0 pffiffi0ffi
4 5
½Q0 Xx ¼ 0 cos60° sin60° ¼ 0 p4 1=2 3=2 5 (f)
ffiffiffi
0 sin 60° cos60° 0  3=2 1=2
620 CHAPTER 11 RIGID BODY DYNAMICS

Step 2:
_
Compute the initial quaternion q 0 using Algorithm 11.2. Substituting Eq. (f) into Eq. (11.150) yields
2 3
0 0 0 0:57735
6 0 0:33333 0 0 7
½K ¼ 6
4 0
7
0 0:33333 0 5
0:57735 0 0 0:66667

Rather than finding the dominant eigenvector by means of the power method, as we did in Example 11.23, we shall here
use MATLAB’s eig function, which obtains all of the eigenpairs. The snippet of MATLAB code for doing so is:
[eigvectors, eigvalues] = eig(K);
%Find the dominant eigenvalue and the column of ‘eigvectors’ that
% contains its eigenvector:
[dominant_eigvalue,column] = max(max(abs(eigvalues)));
dominant_eigenvector = eigvectors(:,column)

The output of this code is,


dominant_eigenvector =
0.5
0
0
0.86603
Therefore, the initial value of the quaternion is

0.5
q0 0 (g)
0
0.86603

Step 3:   T
Specify the initial values of the body frame components of angular velocity ω0 ¼ ωx Þ0 ωy 0 ωz Þ0 :
Recall that Eq. (11.115) relate these body frame angular velocities to the initial values of the top’s Euler angles and their
rates,

ωx Þ0 ¼ ωp 0 sin θ0 sinψ 0 + ωn Þ0 cosψ 0
 
ωy 0 ¼ ωp 0 sin θ0 cosψ 0  ωn Þ0 sinψ 0 (h)

ωz Þ0 ¼ ωs Þ0 + ωp 0 cosθ0

The top is released from rest with a given tilt angle θ0 and spin rate ωs)0. According to Example 11.15,
θ0 ¼ 60°
ψ0 ¼ 0
ωs Þ0 ¼ 1000rpm ¼ 104:72 rad=s (i)

ωp 0 ¼ 51:93rpm ¼ 5:438 rad=s ðlow energy precession rateÞ
ωn Þ0 ¼ 0
Substituting these into Eq. (h) we find
ω0 ¼ b 0 4:7095 107:44 cT ðrad=sÞ (j)
11.11 QUATERNIONS 621

Step 4:
_
Supply ω0 and q 0 as initial conditions to, say, the Runge–Kutta–Fehlberg 4(5) numerical integration procedure (Algorithm
1.3) to solve the system fy_ g ¼ f f g, where
 T
fyg ¼ ωx ωy ωz q1 q2 q3 q4
 T (k)
f f g ¼ dωx =dt dωy =dt dωz =dt dq1 =dt dq2 =dt dq3 =dt dq4 =dt
_
thereby obtaining the angular velocity ω and quaternion q as functions of time. At each step of the numerical integration
process:
_
(i) Use the current value of q to compute [Q]Xx from Algorithm 11.1.
(ii) Use the current value of [Q]Xx and ω to compute dω=dt from (a), (d), and (e).
_ _
(iii) Use the current value of q and ω to compute dq =dt from Eqs. (11.164).
Step 5:
At each solution time:
(i) Use Algorithm 11.1 to compute the direction cosine matrix [Q]Xx.
(ii) Use Algorithm 4.3 to compute the Euler angles ϕ (precession), θ (nutation), and ψ(spin).
(iii) Use Eq. (11.116) to compute the Euler angle rates ϕ, _ θ,
_ and ψ_ .
Step 6:
Plot the time histories of the Euler angles and their rates.
Fig. 11.34 shows the numerical solution for the precession, nutation, and spin angles as well as their rates as functions of
time. We see that the constant precession rate (51.93 rpm) and spin rate (1000 rpm) are in agreement with Example 11.15,

FIG. 11.34
Precession (ϕ), nutation (θ), and spin (ψ) angles and their rates for the top in Example 11.15.
A ¼ B ¼ 0.0012 kg  m2, C ¼ 0.00045 kg  m2.
622 CHAPTER 11 RIGID BODY DYNAMICS

as is the nutation angle, which is fixed at 60°. The sawtooth appearance of the spin angle ψ(t) reflects the fact that it is
confined to the range 0 to 360°.
Clearly , the solution of the steady-state spinning top problem by numerical integration yields no new insight into the
top’s motion and may be deemed a waste of effort. However, suppose we solve the same problem, but release the top from
rest with zero precession rate, so that instead of Eqs. (i), the initial conditions are
θ0 ¼ 60°
ψ0 ¼ 0
ωs Þ0 ¼ 1000rpm ¼ 104:72 rad=s (l)

ωp 0
¼0
ωn Þ0 ¼ 0
_
The initial orientation of the top is unchanged, so the initial quaternion q 0 remains as shown in Eq. (g). On the other hand,
the initially zero precession rate yields a different initial angular velocity vector, namely,

ω0 ¼ b 0 0 104:72 cT ðrad=sÞ (m)


With only this change, the above numerical integration procedure yields the results shown in Fig. 11.35.

FIG. 11.35
Precession (ϕ), nutation (θ), and spin (ψ) angles and their rates for the top in Fig. 11.19, released from rest with
initially zero precession. A ¼ B ¼ 0.0012 kg  m2, C ¼ 0.00045 kg  m2.
PROBLEMS 623

This is an example of unsteady precession, in which we see that the spin axis, instead of making a constant angle of 60° to
the vertical, nutates between 60° and 75.4° at a rate of about 5.7 Hz, while the spin rate itself varies between 975 and
1000 rpm at the same frequency. The precession rate oscillates between 0 and 99.4 rpm, also at a frequency of 5.7 Hz,
with an average rate of 51.9 rpm, which happens to be the steady-state precession rate (Fig. 11.34).
These numerical results can be compared with formulas from the classical analysis of tops in unsteady precession. For
example, it can be shown (Greenwood, 1988) that the relationship between the minimum and maximum nutation angles is
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
cosθmax ¼ λ  λ2  2λ cosθmin + 1
where λ ¼ C2 ωz 2 =ð4AmgdÞ. For the data of this problem, λ ¼ 1.887, so that
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
cosθmax ¼ 1:887  1:8872  2  1:887  cos60° + 1 ¼ 0:2518
θmax ¼ 75:41°
This is precisely what we observe for the nutation angle in Fig. 11.35. By the way, ωz remains constant at its initial value of
104.72 rad/s because the top is axisymmetric (A ¼ B) and Mz ¼ 0 (Eq. (e)3), so that dωz/dt ¼ 0 (Eq. (a)3).

PROBLEMS
Section 2
11.1 Rigid, bent shaft 1 (ABC) rotates at a constant angular velocity of 2K^ rad=s around the positive Z
axis of the inertial frame. Bent shaft 2 (CDE) rotates around BC with a constant angular velocity
of 3^j rad=s, relative to BC. Spinner 3 at E rotates around DE with a constant angular velocity of
4^j rad=s relative to DE. Calculate the magnitude of the absolute angular acceleration vector α3 of
the spinner at the instant shown.
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

{Ans.: kα3 k ¼ 180 + 64sin 2 θ  144 cos θðrad=s2 Þ}

11.2 All the spin rates shown are constant. Calculate the magnitude of the absolute angular
acceleration vector α3 of the spinner at the instant shown (i.e., at the instant when the unit vector ^i
is parallel to the X axis and the unit vector ^j is parallel to the Y axis).
{Ans.: kα3 k ¼ 63 rad=s2 }
624 CHAPTER 11 RIGID BODY DYNAMICS

11.3 The body-fixed xyz frame is attached to the cylinder as shown. The cylinder rotates around the
inertial Z axis, which is collinear with the z axis, with a constant absolute angular velocity θ_ k. ^
Rod AB is attached to the cylinder and aligned with the y axis. Rod BC is perpendicular to AB and
rotates around AB with the constant angular velocity ϕ_ ^j relative to the cylinder. Rod CD is
perpendicular to BC and rotates around BC with the constant angular velocity ν_ m ^ relative to BC,
where m ^ is the unit vector in the direction of BC. The plate abcd rotates around CD with a
^ relative to CD, where the unit vector n
constant angular velocity ψ_ n ^ points in the direction of CD.
^ + ϕ_ ^j + ν_ m
Thus, the absolute angular velocity of the plate is ωplate ¼ θ_ k ^ + ψ_ n
^. Show that
   
(a) ωplate ¼ ðν_ sin ϕ  ψ_ cos ϕsin νÞ^i + ϕ_ + ψ_ cos ν ^j + θ_ + ν_ cos ϕ + ψ_ sinϕ sinν k^
dωplate  
(b) αplate ¼ ¼ ν_ ϕ_ cos ϕ  ψ_ cos ϕcos ν + ψ_ ϕ_ sinϕ sinν  ψ_ θ_ cos ν  ϕ_ θ_ ^i
dt
 
+ ν_ θ_ sinϕ  ψ_ sin ν  ψ_ θ_ cosϕ sinν ^j
 
+ ψ_ ν_ cos ν sinϕ + ψ_ ϕ_ cosϕ sinν  ϕ_ ν_ sin φ_ k^
 
 2  5
(c) aC ¼ l ϕ_ + θ_ 2 sin ϕ^i + 2lϕ_ θ_ cosϕ  lθ_ 2 ^j  lϕ_ 2 cos ϕk ^
4
PROBLEMS 625

11.4 The mass center G of a rigid body has a velocity v ¼ t3^i + 4^j ðm=sÞ and an angular velocity
ω ¼ 2t2 k^ ðrad=sÞ, where t is time in seconds. The ^i, ^j, k
^ unit vectors are attached to and rotate
with the rigid body. Calculate the magnitude of the acceleration aG of the center of mass at
t ¼ 2 s.
{Ans.: aG ¼ 20^i + 64^j ðm=s2 Þ}

11.5 A rigid body is in pure rotation with angular velocity ω ¼ ωx^i + ωy^j + ωz k
^ about the origin of the
^ ^ ^
inertial xyz frame. If point A with position vector rA ¼ 2i + 2j  2k ðmÞ has velocity
vA ¼ ^i + 2^j + 3k
^ ðm=sÞ, what is the magnitude of the velocity of the point B with position vector
^ ^ ^
rB ¼ i + j  k ðmÞ?
{Ans.: 1.871 m/s}
11.6 The inertial angular velocity of a rigid body is ω ¼ ωx^i + ωy^j + ωz k,
^ where ^i, ^j, and k
^ are the unit
^ ^
vectors of a comoving frame whose inertial angular velocity is ω ¼ ωx i + ωy j. Calculate the
components of angular acceleration of the rigid body in the moving frame, assuming that ωx, ωy,
and ωz are all constant.
{Ans.: α ¼ ωy ωz^i  ωx ωz^j}
Section 5
11.7 Find the moments of inertia about the center of mass of the system of six point masses listed in
the table.

Point, i Mass, mi (kg) xi (m) yi (m) zi (m)

1 10 1 1 1
2 10 1 1 1
3 8 4 4 4
4 8 2 2 2
5 12 3 3 3
6 12 3 3 3
2 3
783:5 351:7 40:27
{Ans.: ½IG  ¼ 4 351:7 783:5 80:27 5ðkg  m2 Þg
40:27 80:27 783:5
11.8 Find the mass moment of inertia of the configuration of Problem 11.7 about an axis through
the origin and the point with coordinates (1, 2, 2 m).
{Ans.: 898.7 kg  m2}
626 CHAPTER 11 RIGID BODY DYNAMICS

11.9 A uniform slender rod of mass m and length l lies in the xy plane inclined to the x axis by an
angle θ. Use the results of Example 11.10 to find the mass moments of inertia about the xyz
axes passing through the center of mass G.
2 3
1
6 sin θ  2 sin 2θ 0 7
2

26 7
{Ans.: ½IG  ¼ 12 ml 6 1
1
7}
4  sin 2θ cos 2 θ 0 5
2
0 0 1

11.10 The uniform rectangular box has a mass of 1000 kg. The dimensions of its edges are shown.
(a) Find the mass moments of inertia about the xyz axes.
2 3
1666:7 1500 750
{Ans.: ½IO  ¼ 4 1500 3333:3 500 5ðkg  m2 Þ}
750 500 4333:3
(b) Find the principal moments of inertia and the principal directions about the xyz axes
through O.
{Partial Ans.: I1 ¼ 568:9 kg  m2 , ^e1 ¼ 0:8366^i + 0:4960^j + 0:2326k}
^
(c) Find the moment of inertia about the line through O and the point with coordinates (3 m,
2 m, 1 m).
{Ans.: 583.3 kg  m2}
PROBLEMS 627

11.11 A taxiing airplane turns about its vertical axis with an angular velocity Ω while its propeller
_ Determine the components of the angular momentum of the
spins at an angular velocity ω ¼ θ.
propeller about the body-fixed xyz axes centered at P. Treat the propeller as a uniform slender
rod of mass m and length l.  
{Ans.: Hp ¼ 1 mωl2^i  1 mΩl2 sin 2θ^j + 1 ml2 cos 2 θ + md2 Ωk}
12 24 12
^

11.12 Relative to an xyz frame of reference the components of angular momentum H are given by
2 38 9
1000 0 300 < ωx =  
fHg ¼ 4 0 1000 500 5 ωy kg  m2 =s
: ;
300 500 1000 ωz

where ωx, ωy, and ωz are the components of the angular velocity vector ω. Find the
components ω such that fHg ¼ 1000fωg, where the magnitude of ω is 20 rad/s.
{Ans.: ω ¼ 174:15^i + 10:29^j ðrad=sÞ} 2 3
10 0 0
11.13 Relative to a body-fixed xyz frame ½IG  ¼ 4 0 20 0 5ðkg  m2 Þ and
0 0 30
ω ¼ 2t2^i + 4^j + 3tk
^ ðrad=sÞ, where t is the time in seconds. Calculate the magnitude of the net
moment about the center of mass G at t ¼ 3 s.
{Ans.: 3374 N m}
11.14 In Example 11.11, the system is at rest when a 100-N force is applied to point A as shown.
Calculate the inertial components of angular acceleration at that instant.
{Ans.: αX ¼ 143.9 rad/s2, αY ¼ 553.1 rad/s2, αZ ¼ 7.61 rad/s2}
628 CHAPTER 11 RIGID BODY DYNAMICS

11.15 The body-fixed xyz axes pass through the center of mass G of the airplane and are the principal
axes of inertia. The moments of inertia about these axes are A, B, and C, respectively. The
airplane is in a level turn of radius R with a speed v.
(a) Calculate the bank angle θ.
(b) Use the Euler equations to calculate the rolling moment My that must be applied by the
aerodynamic surfaces.
{Ans.: (a) θ ¼ tan1v2/Rg; (b) My ¼ v2 sin 2θ(C  A)/2R2}

11.16 The airplane in Problem 11.15 is spinning with an angular velocity ωZ about the vertical Z axis.
The nose is pitched down at the angle α. What external moments must accompany this
maneuver?
{Ans.: My ¼ Mz ¼ 0, Mx ¼ ωZ 2 sin 2αðC  BÞ=2}

11.17 Two identical slender rods of mass m and length l are rigidly joined together at an angle θ at
point C, their 2/3 point. Determine the bearing reactions at A and B if the shaft rotates at a
constant angular velocity ω. Neglect gravity and assume that the only bearing forces are normal
to rod AB.
{Ans.: kFAk ¼ mω2l sin θ(1 + 2 cos θ)/18, kFBk ¼ mω2l sin θ(1  cos θ)/9}
PROBLEMS 629

11.18 The flywheel (A ¼ B ¼ 5 kg  m2, C ¼ 10 kg  m2) spins at a constant angular velocity of


^ ðrad=sÞ. It is supported by a massless gimbal that is mounted on the platform as
ωs ¼ 100k
shown. The gimbal is initially stationary relative to the platform, which rotates with a constant
angular velocity of ωp ¼ 0:5^j ðrad=sÞ. What will be the gimbal’s angular acceleration when the
torquer applies a torque of 600^i ðN mÞ to the flywheel?
{Ans.: 70^i ðrad=s2 Þ}

11.19 A uniform slender rod of length L and mass m is attached by a smooth pin at O to a vertical shaft
that rotates at constant angular velocity ω. Use the Euler equations and the body frame shown to
calculate ω p at ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
the instant shown.
{Ans.: ω ¼ 3g=ð2L cos θÞ}
630 CHAPTER 11 RIGID BODY DYNAMICS

11.20 A uniform, thin circular disk of mass 10 kg spins at a constant angular velocity of 630 rad/s
about axis OG, which is normal to the disk and pivots about the frictionless ball joint at O.
Neglecting the mass of the shaft OG, determine the rate of precession if OG remains horizontal
as shown. Gravity acts down, as shown. G is the center of mass and the y axis remains fixed in
space. The moments of inertia about G are IG)z ¼ 0.02812 kg  m2 and
IG)x ¼ IG)y ¼ 0.01406 kg  m2.
{Ans.: 1.38 rad/s}

Section 7
11.21 Consider a rigid body experiencing rotational motion associated with an angular velocity
vector ω. The inertia tensor (relative to body-fixed axes through the center of mass G) is
2 3
20 10 0  
4 10 30 0 5 kg  m2
0 0 40

and ω ¼ 10^i + 20^j + 30k


^ ðrad=sÞ. Calculate
(a) the angular momentum HG and
(b) the rotational kinetic energy (about G).
{Partial Ans.: (b) TR ¼ 23, 000 J}
PROBLEMS 631

Section 8
11.22 At the end of its takeoff run, an airplane with retractable landing gear leaves the runway
with a speed of 130 km/h. The gear rotates into the wing with an angular velocity of 0.8 rad/
s with the wheels still spinning. Calculate the gyroscopic bending moment in the wheel
bearing B. The wheels have a diameter of 0.6 m, a mass of 25 kg, and a radius of gyration of
0.2 m.
{Ans.: 96.3 N m}

11.23 The gyro rotor, including shaft AB, has a mass of 4 kg and a radius of gyration 7 cm around AB.
The rotor spins at 10,000 rpm while also being forced to rotate around the gimbal axis CC at
2 rad/s. What are the transverse forces exerted on the shaft at A and B? Neglect gravity.
{Ans.: 1.03 kN}

11.24 A jet aircraft is making a level, 2.5-km radius turn to the left at a speed of 650 km/h. The rotor of
the turbojet engine has a mass of 200 kg, a radius of gyration of 0.25 m, and rotates at
15,000 rpm clockwise as viewed from the front of the airplane. Calculate the gyroscopic
moment that the engine exerts on the airframe and explain why it tends to pitch the nose up
or down.
{Ans.: 1.418 kN m; pitch down}

11.25 A cylindrical rotor of mass 10 kg, radius 0.05 m, and length 0.60 m is simply supported at each
end in a cradle that rotates at a constant 20 rad/s counterclockwise as viewed from above.
Relative to the cradle, the rotor spins at 200 rad/s counterclockwise as viewed from the right
632 CHAPTER 11 RIGID BODY DYNAMICS

(from B toward A). Assuming that there is no gravity, calculate the bearing reactions RA and RB.
Use the comoving xyz frame shown, which is attached to the cradle but not to the rotor.
{Ans.: RA ¼  RB ¼ 83.3 N}

Section 9
11.26 The Euler angles of a rigid body are ϕ ¼ 50°, θ ¼ 25°, and ψ ¼ 70°. Calculate the angle (a
positive number) between the body-fixed x axis and the inertial X axis.
{Ans.: 115.6°}
Section 11
11.27 Let ^i^jk
^ and ^I^
JK^ be two right-handed triads of orthogonal unit vectors related as in Eq. (4.18) by
the direction cosine matrix [Q], so that

^i ¼ Q11^I + Q12 ^ ^
J + Q13 K
^j ¼ Q21^I + Q22 ^ ^
J + Q23 K
^ ¼ Q31^I + Q32 ^
k ^
J + Q33 K
Show that ^i ¼ ^j  k
^ implies that Q13 ¼ Q32Q21  Q22Q31, whereas ^j ¼ k
^ ^i implies that
Q23 ¼ Q12 Q31  Q32 Q11 :

REFERENCES
Bar-Itzhack, I.Y., 2000. New Method for Extracting the Quaternion From the Rotation Matrix. J. Guid. Control
Dyn. 23, 1085–1087.
Jennings, A., 1977. Matrix Computation for Engineers and Scientists. Wiley, New York.
Greenwood, D.T., 1988. Principles of Dynamics, second ed. Prentice-Hall, Englewood Cliffs, NJ.
Gelman, H., 1971. A note on the time dependence of the effective axis and angle of a rotation. J. Res. Natl. Bureau
Stand. B Math. Sci. 72B (3 and 4), 168.
Hughes, P.C., 2004. Spacecraft Attitude Dynamics. Dover p. 25.

You might also like