You are on page 1of 358

S.

Widnall
16.07 Dynamics
Fall 2009
Version 2.0

Lecture L1 - Introduction

Introduction
In this course we will study Classical Mechanics and its application to aerospace systems. Particle motion in
Classical Mechanics is governed by Newtons laws and is sometimes referred to as Newtonian Mechanics. The
motion of extended rigid bodies is analyzed by application of Newtons law to a multi-particle system. These
laws are empirical in that they combine observations from nature and some intuitive concepts. Newtons laws
of motion are not self evident. For instance, in Aristotelian mechanics before Newton, a force was thought
to be required in order to maintain motion. Much of the foundation for Newtonian mechanics was laid by
Galileo at the end of the 16th century. Newton, in the middle of the 17th century stated the laws of motion
in the form we know and use them today, and shortly after, he formulated the law of universal attraction.
This led to a complete theory with which he was able to explain many observed phenomena, in particular
the motion of the planets. Nevertheless, these laws still left many unanswered questions at that time, and
it was not until later years that the principles of classical mechanics were deeply studied and rationalized.
In the eighteenth century, there were many contributions in this direction, such as the principle of virtual
work by Bernoulli, DAlamberts principle and the theory of rigid body dynamics developed by Euler. In the
nineteenth century, Lagrange and later Poisson, Hamilton and Jacobi developed the so called analytical or
rational mechanics and gave to the theory of Newtonian mechanics a much richer mathematical structure.
Classical Mechanics has its limitations and breaks down where more modern theories such as relativity and
quantum mechanics, developed in the twentieth century, are successful. Newtonian mechanics breaks down
for systems moving at speeds comparable with the speed of light, and also fails for systems of dimensions
comparable to the size of the atom. Nevertheless, for practical engineering applications, Newtonian mechanics
provides a very good model to represent reality, and, in fact, it is hard to nd examples in aerospace where
Newtonian mechanics is not adequate. The most notable perhaps are the relativistic corrections that need
to be made for modeling satellite communications.

16.07s Place in the Aero-Astro Curriculum


Aero-Astro focuses on the analysis, design and control of aerospace vehicles, both aircraft and space craft
and the environment in which they are used. The place of 16.07 within the overall curriculum is shown in
Figure 1.
1

16.07 is a core discipline of aerospace engineering: dealing with the natural dynamics of aero-astro systems.
The complexities of aerodynamic forces (as seen in unied engineering and in 16.100) or structural exibility
(as covered in unied engineering and in 16.20) and their eect on vehicle dynamics are treated very simply,
if at all. Beyond the study of the natural dynamics of aircraft and space craft, with or without aerodynamic
forces and structural exibility, is our need to develop approaches to control the behavior of the system.
Thus 16.06-Automatic Control surrounds this group of courses, moving beyond the natural dynamics to
impose control laws upon the system.

Image removed due to copyright restrictions.

Depending upon our interests, 1) in the dynamic motion of aircraft, where we would want to predict
2

position, velocities, and acceleration under the action of forces and moments as well as aircraft stability

Image removed due to copyright restrictions.

or 2) in the motion of spacecraft, orbits, satellite stability, launch dynamics, and orbit transfers,

Image removed due to copyright restrictions.

we will model the system as simply as possible for our purposes. If we are interested in the earths motion
about the sun, we will model both the sun and the earth as point masses of no extent. To determine the

motion of a satellite in orbit, we may model the satellite as a point mass, or if we are concerned whether the
satellite will tip over, we will model it as a body of nite extent.

Particles, Rigid Bodies and Real Bodies


In this course real bodies will be idealized either as particles or as rigid bodies.

A particle is a body of negligible dimensions. When the dimensions of the body are unimportant to the
description of its motion, we will idealize the body as a particle.

A rigid body is a body that has a nite size but does not deform. This will be a useful approximation when
the deformation of a body is negligible compared to the overall motion. For instance, we may consider an
aircraft as being a rigid body when considering the behavior of the aircraft along its ight path, even though
under some specic conditions the deection of the wing tips may be considerable. In describing the motion
of a rigid body, we need to be concerned not only with its position but also with its orientation.

On the other hand, real bodies have a nite size and are always deformable under loading. In some situations
it will be required to consider the deformation of the body when considering its dynamic behavior, but this
is outside the scope of this course, except perhaps when we introduce the topics of vibration and waves.

Scalars and Vectors


Scalars and vectors are mathematical abstractions that are very useful to describe many of the concepts in
dynamics. You should already have an intuitive idea of what they are. A scalar is a single number which is
useful to describe the reading of a physical property on a scale. For instance, temperature, length or speed
are scalar quantities. On the other hand, vectors are much richer entities. They exist in a multi-dimensional
space and they have both direction and magnitude. Velocities, forces and electric elds are examples of
vector quantities.

Newtons Laws
Newtons three laws of motion are:
1.- A particle in isolation moves with constant velocity.
A particle in isolation means that the particle does not interact with any other particle. Constant
velocity means that the particle moves along a straight line with constant speed. In particular, it
can be at rest. It turns out that the motion (e.g. velocity and acceleration) we observe depends on
the reference frame we use. Therefore, the above law cannot be veried in all reference frames. The

reference frames for which this law is satised are called inertial reference frames. In some sense, we
can say that Newtons rst law postulates that inertial reference frames exist.
2.- The acceleration of a particle relative to an inertial reference frame is equal to the force per unit mass
applied to the particle.
In other words, if F represents the (vector) sum of all forces acting on a particle of mass m, any inertial
observer will see that the particle has an acceleration a which is given by,
F = ma .

(1)

This equation introduces two new concepts: force and mass. Precise denitions for these concepts are
not easy even though we all have some intuition about both force and mass. Forces result when bodies
interact. Forces are vectors with magnitude and direction. They are measured by comparison, e.g. to
the weight of a standard mass, or to the deformation of a spring. We will assume that the concept
of force is absolute and does not depend on the observer. Once we have dened force, we can dene
mass as the constant of proportionality between force and acceleration. Mass is a scalar quantity. One
could think of mass as the resistance of bodies to a change in motion. That is, a given force applied to
a body with small mass will produce a large acceleration, whereas the same force on a body of large
mass will produce a small acceleration.
Equation (1) is a vector equation. This means that the force and the acceleration always have the
same direction and the ratio of their magnitudes is m. It turns out that this equation is the basis for
most engineering dynamics.
3.- The forces of action and reaction between interacting bodies are equal in magnitude and opposite in
direction
This law makes explicit the fact that a force is the result of interaction between bodies. This law is
clearly satised when the bodies are in contact and in static equilibrium. The situation for bodies
in motion interacting at a distance, e.g. electromagnetic or gravity interactions, is a little bit more
complicated. We know that electromagnetic signals travel at a nite speed and therefore there is a
time delay whenever two bodies interact at a distance. Unfortunately, Newton did not foresee such a
situation, and in these cases, Newtons third law breaks down. However, for practical purposes and for
most engineering applications, the error made by assuming that these interactions are instantaneous,
and, hence, assuming that Newtons third law is applicable, is negligible.

Note

Units

We shall primarily use two systems of units: the International System, also called SI, and the English System.

The international system is the most widely used system for science and engineering. The English system,

however, is still in widespread use within the United States engineering community. For this reason we will
use and should feel comfortable with both systems. The SI units are meter (m), kilogram (kg), and second
(s) for length, mass and time, respectively. Acceleration is measured in m/s2 , and force is measured in kgm
/s2 , which is also called a Newton (N) i.e. 1N = 1Kgm/s2 . In the English system, the units of length,
mass and time are foot (ft), slug, and second (s), respectively. Acceleration is measured in ft/s2 , and force is
measured in slug ft/s2 , also called pound (lb), i.e. 1 pound = 1slugft/s2 . We have the following conversion
factors:
1 ft

0.348 m

1 slug

14.5 kg

1N

0.224 lb

Example

Inertial vs. Non-inertial observers

This example is meant to illustrate the fact that we can easily come up with situations in which Newtons
second law is not satised for accelerating, non-inertial observers. We will come back to this example later
on in the course.

Consider a rocket sled which can move on a horizontal track as shown. We consider an inertial observer
O which is xed on the ground and an observer O which is on the sled. We also have an accelerometer
mounted on the sled. This consists of a known proof mass m, whose horizontal motion relative to the sled
is constrained by a spring. We assume that the friction between the mass and the sled is negligible. This
means that the only mechanism to exert a horizontal force on the mass is through the spring.

We consider two situations:


The engine is o, T = 0, a = 0, the mass is at rest and the spring is uncompressed. Both observers O
and O agree in all their measurements.
The engine is on, T is a constant non-zero force, the sled has an acceleration a, and the spring is
compressed and exerting a force F on the mass. Both observers are able to measure the force exerted
by the spring by reading on a scale how much the spring deforms. For observer O , the mass is not

moving (assume that the initial transient oscillation is gone and the spring has settled to a xed position
relative to the sled). For the observer on the ground the mass is accelerating with acceleration a. Since
observer O is an inertial observer, he or she is able to verify that F = ma. On the other hand, the
observer on the sled, O , measures a force F , but observes a zero acceleration. Hence, the observations
of observer O do not satisfy Newtons second law.

Law of Universal Attraction


The law of universal attraction was proposed by Newton shortly after formulating the laws of motion. The
law postulates that the force of attraction between any two particles, of masses M and m, has a magnitude,
F , given by
F =G

Mm
r2

(2)

where r is the distance between the two particles, and G = 6.673(1011 ) m3 /(kg s2 ) is the universal constant
of gravitation which is determined according to experimental evidence. The direction of the force is parallel
to the line connecting the two particles.

The law of gravitation stated above is strictly valid for point masses. One would expect that when the size
of the masses is comparable to the distance between the masses one would observe deviations to the above
law. It turns out that if the mass M is distributed uniformly over a sphere of radius R, the force on a mass
m, outside M , is still given by (2), with r being measured from the spheres center.

Weight
The gravitational attraction from the earth to any particle located near the surface of the earth is called the
weight. Thus, the weight, W, of a particle of mass m at sea level is given by
W = G

Me m
er = g0 mer = mg 0 .
Re2

Here, Me 5.976 1024 kg and Re 6.371 106 m, are the mass and radius of the earth, respectively, and
g 0 = (GMe /Re2 ) er , is the gravitational acceleration vector at sea level. The average value of its magnitude
is g0 = 9.825 m/s2 .
The variation of the gravitational attraction with altitude is easily determined from the gravitational law.
Thus, the weight at an altitude h above sea level is given by
W = G

Me m
Re2
Re2
e
=
g
me
=
m
g .
r
0
r
(Re + h)2
(Re + h)2
(Re + h)2 0

It turns out that the earth is not quite spherical and so the weight does not exactly obey the inverse-squared
law. The magnitude of the gravitational acceleration, g0 , at the poles and at the equator, is slightly dierent.
In addition, the earth is also rotating. As we shall see this introduces an inertial centrifugal force which has
the eect of reducing the vertical component of the weight. We will study these eects later on in the course.

References
[1] R. Dugas, A History of Mechanics, Dover, 1988.
ADDITIONAL READING
J.L. Meriam and L.G. Kraige, Engineering Mechanics, DYNAMICS, 5th Edition
1/1, 1/2, 1/3, 1/4, 1/5 (Eect or Altitude only), 1/6, 1/7, 3/2

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

S. Widnall
16.07 Dynamics
Fall 2009
Version 1.0

Lecture L2 - Degrees of Freedom and Constraints, Rectilinear

Motion

Degrees of Freedom
Degrees of freedom refers to the number of independent spatial coordinates that must be specied to deter
mine the position of a body. If the body is a point mass, only three coordinates are required to determine
its position. On the other hand, if the body is extended, such as an aircraft, three position coordinates and
three angular coordinates are required to completely specify its position and orientation in space.

Kinematic Constraints
In many situations the number of independent coordinates will be reduced below this number, either because
the number of spacial dimensions is reduced or because there are relationships specied among the spatial
coordinates. When setting up problems for solution it is useful to think of these relationships as constraints.

For example, if a point mass is constrained to move in a plane (two dimensions) the number of spatial
coordinates necessary to describe its motion is two. If instead of being a point mass, this body has extended
dimensions, such as a at plate conned to a plane, it requires three coordinates to specify its position and
orientation: two position coordinates and one angular coordinate.

If a particle is conned to move on a curve in either two or three dimensions, such as a bead moving on a
wire, the number of independent coordinates necessary to describe its motion is one.

Another source of constraints on the motion of particles is connections between them. For example, the two
particle connected by a cable passing over a pulley are constrained to move in equal and opposite directions.
More complex arrangements are possible and can be analyzed using these ideas. Two gears in contact are
constrain to move together according to their individual geometry.

A cylinder rolling on a plane is constrained in two ways. Contact with the plane reduces the two-dimensional
motion to one spatial coordinate along the plane, and the constraint of rolling provides a relationship be
tween the angular coordinates and the spatial position, resulting in a single degree of freedom system.

Internal Force-Balance Constraints


Another type of constraint occurs when we consider the of a system of particles and the necessary force
balance that occurs between the parts. These constraints follow directly from Newtons third law: the force
of action and reaction between two bodies are equal in magnitude and opposite in direction. We will pursue
these ideas in greater depth later in the course. For now, we will give a simple example to illustrate the
principle.
Consider the systems shown in a) and b).

System a) consists of two masses m in contact resting on a frictionless plane in the presence of gravity. A
force F is applied to mass 1, and it is obvious that the two masses will accelerate at a = F/(2m). If we
look at the two masses separately, we can determine what internal force must exist between them to cause
the motion. It is clear that each mass feels a net force of F/2, since its acceleration is a = F/(2m). This
net force arises because between the two masses there is an equal and opposite force F/2 acting across the
interface. Another way to look at this is that the interface between the bodies is a body of zero mass, and
therefore can have no net force acting upon it otherwise its acceleration would be innite.
System b) is a bit more complex, primarily because the forces between one mass and the mass above it
are shear forces and must be supplied by friction. Assuming that the friction coecient is large enough to
accelerate the three masses an equal amount given by a = F/(3m), by the reasoning we have discussed, the
force balance is as sketched in b): equal and opposite normal forces F 2/3 on the vertical surfaces, and
equal and opposite shear forces F 1/3 on the horizontal surfaces.

Rectilinear Motion
In many case we can get an exact expression for the position of a particle as a function of time. We start by
considering the simple motion of a particle along a straight line. The position of particle A at any instant
can be specied by the coordinate s with origin at some xed point O.

The instantaneous velocity is


v=

ds
= v .
dt

(1)

We will be using the dot notation, to indicate time derivative, e.g. () d/dt. Here, a positive v means
that the particle is moving in the direction of increasing s, whereas a negative v, indicates that the particle
is moving in the opposite direction. The acceleration is
a=

dv
d2 s
= v = 2 = s .
dt
dt

(2)

The above expression allows us to calculate the speed and the acceleration if s and/or v are given as a
function of t, i.e. s(t) and v(t). In most cases however, we will know the acceleration and then, the velocity
and the position will have to be determined from the above expressions by integration.

Determining the velocity from the acceleration


From a(t)
If the acceleration is given as a function of t, a(t), then the velocity can be determined by simple integration
of equation (2),
t

v(t) = v0 +

a(t) dt .

(3)

t0

Here, v0 is the velocity at time t0 , which is determined by the initial conditions.


From a(v)
If the acceleration is given as a function of velocity a(v), then, we can still use equation (2), but in this case
we will solve for the time as a function of velocity,

t(v) = t0 +
v0

dv
.
a(v)

(4)

Once the relationship t(v) has been obtained, we can, in principle, solve for the velocity to obtain v(t).
A typical example in which the acceleration is known as a function of velocity is when aerodynamic drag
forces are present. Drag forces cause an acceleration which opposes the motion and is typically of the form
a(v) v 2 (the sign means proportional to, that is, a(v) = v 2 for some , which is not a function of
velocity).
From a(s)
When the acceleration is given as a function of s then, we need to use a combination of equations (1) and
(2), to solve the problem. From
a=

dv
dv ds
dv

=
=v
dt
ds dt
ds
4

(5)

we can write
a ds = v dv .
This equation can now be used to determine v as a function of s,
s
v 2 (s) = v02 + 2
a(s) ds .

(6)

(7)

s0

where, v0 , is the velocity of the particle at point s0 . Here, we have used the fact that,
v
v
v2
v2
v2
0 .
v dv =
d( ) =
2
2
2
v0
v0
A classical example of an acceleration dependent on the spatial coordinate s, is that induced by a deformed
linear spring. In this case, the acceleration is of the form a(s) s.
Of course, when the acceleration is constant, any of the above expressions (3, 4, 7), can be employed. In
this case we obtain,
v = v0 + a(t t0 ),

or

v 2 = v02 + 2a(s s0 ) .

or

v 2 = v02 + 2g(s s0 ) .

If a = g, this reduces to the familiar

v = v0 + g(t t0 ),

Determining the position from the velocity


Once we know the velocity, the position can be found by integrating ds = vdt from equation (1). Thus,
when the velocity is known as a function of time we have,
t
s = s0 +
v(t) dt .

(8)

t0

If the velocity is known as a function of position, then

t = t0 +

s0

ds
.
v(s)

(9)

Here, s0 is the position at time t0 .


It is worth pointing out that equation (6), can also be used to derive an expression for v(s), given a(v),
s
v
v
s s0 =
ds =
dv .
(10)
a(v)
s0
v0
This equation can be used whenever equation (4) is applicable and gives v(s) instead of t(v). For the case
of constant acceleration, either of equations (8, 9), can be used to obtain,
1
s = s0 + v0 (t t0 ) + a(t t0 )2 .
2

In many practical situations, it may not be possible to carry out the above integrations analytically in which
case, numerical integration is required. Usually, numerical integration will also be required when either the
velocity or the acceleration depend on more than one variable, i.e. v(s, t), or, a(s, v).

Example

Reentry, Ballistic Coecient, Terminal Velocity

Terminal Velocity
Terminal velocity occurs when the acceleration becomes zero and the velocity Consider an air-dropped
payload starting from rest. The force on the body is a combination of gravity and air drag and has the form
1
F = mg v 2 CD A
2

(11)

Applying Newtons law and solving for the acceleration a we obtain


1
CD A
a = g v 2
m
2
The quantity

m
CD A

(12)

characterizes the combined eect of body shape and mass on the acceleration; it is an

important parameter in the study of reentry; it is called the Ballistic Coecient. It is dened as =

m
CD A .

Unlike many coecients that appear in aerospace problems, the Ballistic Coecient is not non-dimensional,
but has units of mass/length2 or kg/meters2 in mks units. Also, in some applications, the ballistic coecient
is dened as the inverse, B =

CD A
m ,

so it pays to be careful in its application. Equation 13 then becomes


1
a = g v 2 /
2

(13)

Terminal velocity occurs when the force of gravity equals the drag on the object resulting in zero acceleration.
This balance gives the terminal velocity as

vterminal =

2gm
=
CD A

2g

(14)

For the Earth, atmospheric density at sea level is = 1.225kg/m3 ; we shall deal with the variation of
atmospheric density with altitude when we consider atmospheric reentry of space vehicles. Typical value of
range from = 1 (Assuming CD = .5, a tennis ball has = 35.) to = 1000 for a reentry vehicle. As an
example, consider a typical case of = 225, where the various parameters then give the following expression
for the acceleration
a = g 0.002725v 2 m/s .
Here g = 9.81m/s2 , is the acceleration due to gravity and v is the downward velocity. It is clear from this
expression that initially the acceleration will be g. Therefore, the velocity will start to increase and keep on
increasing until a = 0, at which point the velocity will stay constant. The terminal velocity is then given by,
0 = g 0.002725 vf2

or

vf = 60m/s .

To determine the velocity as a function of time, the acceleration can be re-written introducing the terminal
velocity as, a = g(1 (v/vf )2 ) . We then use expression (4), and write

1 v
dv
1 v
1/2
1/2
vf
vf + v
t=
=
(
+
)d
v
=
ln
.
g 0 1 (v/vf )2
g 0 1 + (v/vf ) 1 (v/vf )
2g
vf v
Solving for v we obtain,
v = vf

e2gt/vf 1
m/s .
e2gt/vf + 1

We can easily verify that for large t, v = vf . We can also nd out how long does it take for the payload to
reach, say, 95% of the terminal velocity,
t=

vf
1.95
ln
= 11.21s .
2g 0.05

To obtain an expression for the velocity as a function of the traveled distance we can use expression (10)
and write
s=

1
g

vf2
v dv
=

ln(1 (v/vf )2 ) .
1 (v/vf )2
2g

Solving for v we obtain

2
v = vf 1 e2gs/vf m/s .
We see that for, say, v = 0.95vf , s = 427.57m. This is the distance traveled by the payload in 11.21s,
which can be compared with the distance that would be traveled in the same time if we were to neglect air
resistance, sno

drag

= gt2 /2 = 615.75m.

Example

Spring-mass system

Here, we consider a mass allowed to move without friction on a horizontal slider and subject to the force
exerted by a linear spring. Initially the system is in equilibrium (no force on the spring) at s = 0. Suddenly,
the mass is given a velocity v0 and then the system is left free to oscillate. We know that the eect of the
spring is to cause an acceleration to the body, opposing the motion, of the form a = s, where > 0 is a
constant.

Using equation (7), we have


v 2 = v02 s2 .
The displacement can now be obtained using expression (9),

s
s
ds
1

arcsin
t=
=
,
2 s2

v
0
v
0
0

which gives,

v0
s = sin t .

Finally, the velocity as a function of time is simply, v = v0 cos

t. We recognize this motion as that of an

undamped harmonic oscillator.

References
[1] R. Dugas, A History of Mechanics, Dover, 1988.
ADDITIONAL READING
J.L. Meriam and L.G. Kraige, Engineering Mechanics, DYNAMICS, 5th Edition
1/1, 1/2, 1/3, 1/4, 1/5 (Eect or Altitude only), 1/6, 1/7, 3/2

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

S. Widnall
16.07 Dynamics
Fall 2009
Lecture notes based on J. Peraire Version 2.0

Lecture L3 - Vectors, Matrices and Coordinate Transformations


By using vectors and dening appropriate operations between them, physical laws can often be written in
a simple form. Since we will making extensive use of vectors in Dynamics, we will summarize some of their
important properties.

Vectors
For our purposes we will think of a vector as a mathematical representation of a physical entity which has
both magnitude and direction in a 3D space. Examples of physical vectors are forces, moments, and velocities.
Geometrically, a vector can be represented as arrows. The length of the arrow represents its magnitude.
Unless indicated otherwise, we shall assume that parallel translation does not change a vector, and we shall
call the vectors satisfying this property, free vectors. Thus, two vectors are equal if and only if they are
parallel, point in the same direction, and have equal length.
Vectors are usually typed in boldface and scalar quantities appear in lightface italic type, e.g. the vector
quantity A has magnitude, or modulus, A = |A|. In handwritten text, vectors are often expressed using the

arrow, or underbar notation, e.g. A , A.

Vector Algebra
Here, we introduce a few useful operations which are dened for free vectors.

Multiplication by a scalar
If we multiply a vector A by a scalar , the result is a vector B = A, which has magnitude B = ||A. The
vector B, is parallel to A and points in the same direction if > 0. For < 0, the vector B is parallel to
A but points in the opposite direction (antiparallel).

If we multiply an arbitrary vector, A, by the inverse of its magnitude, (1/A), we obtain a unit vector which
eA , etc. Thus, we
is parallel to A. There exist several common notations to denote a unit vector, e.g. A,
= A/A = A/|A|, and A = A A,
|A|
= 1.
have that A
1

Vector addition
Vector addition has a very simple geometrical interpretation. To add vector B to vector A, we simply place
the tail of B at the head of A. The sum is a vector C from the tail of A to the head of B. Thus, we write
C = A + B. The same result is obtained if the roles of A are reversed B. That is, C = A + B = B + A.
This commutative property is illustrated below with the parallelogram construction.

Since the result of adding two vectors is also a vector, we can consider the sum of multiple vectors. It can
easily be veried that vector sum has the property of association, that is,
(A + B) + C = A + (B + C).

Vector subtraction
Since A B = A + (B), in order to subtract B from A, we simply multiply B by 1 and then add.

Scalar product (Dot product)


This product involves two vectors and results in a scalar quantity. The scalar product between two vectors
A and B, is denoted by A B, and is dened as
A B = AB cos .
Here , is the angle between the vectors A and B when they are drawn with a common origin.

We note that, since cos = cos(), it makes no dierence which vector is considered rst when measuring
the angle . Hence, A B = B A. If A B = 0, then either A = 0 and/or B = 0, or, A and B are
orthogonal, that is, cos = 0. We also note that A A = A2 . If one of the vectors is a unit vector, say
= A cos , is the projection of vector A along the direction of B
.
B = 1, then A B

Exercise
Using the denition of scalar product, derive the Law of Cosines which says that, for an arbitrary triangle
with sides of length A, B, and C, we have
C 2 = A2 + B 2 2AB cos .
Here, is the angle opposite side C. Hint : associate to each side of the triangle a vector such that C = AB,
and expand C 2 = C C.

Vector product (Cross product)


This product operation involves two vectors A and B, and results in a new vector C = AB. The magnitude
of C is given by,
C = AB sin ,
where is the angle between the vectors A and B when drawn with a common origin. To eliminate ambiguity,

between the two possible choices, is always taken as the angle smaller than . We can easily show that C

is equal to the area enclosed by the parallelogram dened by A and B.

The vector C is orthogonal to both A and B, i.e. it is orthogonal to the plane dened by A and B. The

direction of C is determined by the right-hand rule as shown.

From this denition, it follows that


B A = A B ,
which indicates that vector multiplication is not commutative (but anticommutative). We also note that if
A B = 0, then, either A and/or B are zero, or, A and B are parallel, although not necessarily pointing
in the same direction. Thus, we also have A A = 0.
Having dened vector multiplication, it would appear natural to dene vector division. In particular, we
could say that A divided by B, is a vector C such that A = B C. We see immediately that there are a
number of diculties with this denition. In particular, if A is not perpendicular to B, the vector C does
not exist. Moreover, if A is perpendicular to B then, there are an innite number of vectors that satisfy
A = B C. To see that, let us assume that C satises, A = B C. Then, any vector D = C + B, for

any scalar , also satises A = B D, since B D = B (C + B) = B C = A. We conclude therefore,


that vector division is not a well dened operation.
Exercise
Show that |A B| is the area of the parallelogram dened by the vectors A and B, when drawn with a
common origin.

Triple product
Given three vectors A, B, and C, the triple product is a scalar given by A (B C). Geometrically, the
triple product can be interpreted as the volume of the three dimensional parallelepiped dened by the three
vectors A, B and C.

It can be easily veried that A (B C) = B (C A) = C (A B).

Exercise
Show that A (B C) is the volume of the parallelepiped dened by the vectors A, B, and C, when drawn
with a common origin.

Double vector product


The double vector product results from repetition of the cross product operation. A useful identity here is,
A (B C) = (A C)B (A B)C .
Using this identity we can easily verify that the double cross product is not associative, that is,
A (B C) = (A B) C .

Vector Calculus
Vector dierentiation and integration follow standard rules. Thus if a vector is a function of, say time, then
its derivative with respect to time is also a vector. Similarly the integral of a vector is also a vector.

Derivative of a vector
Consider a vector A(t) which is a function of, say, time. The derivative of A with respect to time is dened
as,
dA
A(t + t) A(t)
= lim
.
t0
dt
t

(1)

A vector has magnitude and direction, and it changes whenever either of them changes. Therefore the rate
of change of a vector will be equal to the sum of the changes due to magnitude and direction.
Rate of change due to magnitude changes
When a vector only changes in magnitude from A to A + dA, the rate of change vector dA is clearly parallel
to the original vector A.

Rate of change due to direction changes


Let us look at the situation where only the direction of the vector changes, while the magnitude stays
constant. This is illustrated in the gure where a vector A undergoes a small rotation. From the sketch, it
is clear that if the magnitude of the vector does not change, dA is perpendicular to A and as a consequence,
the derivative of A, must be perpendicular to A. (Note that in the picture dA has a nite magnitude and
therefore, A and dA are not exactly perpendicular. In reality, dA has innitesimal length and we can see
that when the magnitude of dA tends to zero, A and dA are indeed perpendicular).

An alternative, more mathematical, explanation can be derived by realizing that even if A changes but its
modulus stays constant, then the dot product of A with itself is a constant and its derivative is therefore
zero. A A = constant. Dierentiating, we have that,
dA A + A dA = 2A dA = 0 ,
which shows that A, and dA, must be orthogonal.

Suppose that A is instantaneously rotating in the plane of the paper at a rate = d/dt, with no change in
and the magnitude of dA, will be
magnitude. In an instant dt, A, will rotate an amount d = dt
.
dA = |dA| = Ad = Adt
Hence, the magnitude of the vector derivative is

dA

dt = A .
In the general three dimensional case, the situation is a little bit more complicated because the rotation
of the vector may occur around a general axis. If we express the instantaneous rotation of A in terms of
an angular velocity (recall that the angular velocity vector is aligned with the axis of rotation and the
direction of the rotation is determined by the right hand rule), then the derivative of A with respect to time
is simply,

dA
dt

=A .

(2)

constant magnitude

To see that, consider a vector A rotating about the axis C C with an angular velocity . The derivative
will be the velocity of the tip of A. Its magnitude is given by l, and its direction is both perpendicular to
A and to the axis of rotation. We note that A has the right direction, and the right magnitude since
l = A sin .

Expression (2) is also valid in the more general case where A is rotating about an axis which does not pass
through the origin of A. We will see in the course, that a rotation about an arbitrary axis can always be
written as a rotation about a parallel axis plus a translation, and translations do not aect the magnitude
not the direction of a vector.
We can now go back to the general expression for the derivative of a vector (1) and write

dA
dA
dA
dA
=
+
=
+A .
dt
dt constant direction
dt constant magnitude
dt constant direction
Note that (dA/dt)constant direction is parallel to A and A is orthogonal to A. The gure below shows
the general dierential of a vector, which has a component which is parallel to A, dA , and a component
which is orthogonal to A, dA . The magnitude change is given by dA , and the direction change is given
by dA .
6

Rules for Vector Dierentiation


Vector dierentiation follows similar rules to scalars regarding vector addition, multiplication by a scalar,
and products. In particular we have that, for any vectors A, B, and any scalar ,
d(A)

dA + dA

d(A + B)

dA + dB

d(A B)
d(A B)

= dA B + A dB
=

dA B + A dB .

Components of a Vector
We have seen above that it is possible to dene several operations involving vectors without ever introducing
a reference frame. This is a rather important concept which explains why vectors and vector equations are
so useful to express physical laws, since these, must be obviously independent of any particular frame of
reference.
In practice however, reference frames need to be introduced at some point in order to express, or measure,
the direction and magnitude of vectors, i.e. we can easily measure the direction of a vector by measuring
the angle that the vector makes with the local vertical and the geographic north.
Consider a right-handed set of axes xyz, dened by three mutually orthogonal unit vectors i, j and k
(i j = k) (note that here we are not using the hat () notation). Since the vectors i, j and k are mutually
orthogonal they form a basis. The projections of A along the three xyz axes are the components of A in the
xyz reference frame.

In order to determine the components of A, we can use the scalar product and write,
Ax = A i,

Ay = A j,
7

Az = A k .

The vector A, can thus be written as a sum of the three vectors along the coordinate axis which have
magnitudes Ax , Ay , and Az and using matrix notation, as a column vector containing the component
magnitudes.

Ax

A = Ax + Ay + Az = Ax i + Ay j + Az k = Ay .

Az

Vector operations in component form


The vector operations introduced above can be expressed in terms of the vector components in a rather
straightforward manner. For instance, when we say that A = B, this implies that the projections of A and
B along the xyz axes are the same, and therefore, this is equivalent to three scalar equations e.g. Ax = Bx ,
Ay = By , and Az = Bz . Regarding vector summation, subtraction and multiplication by a scalar, we have
that, if C = A + B, then,
Cx = Ax + Bx ,

Cy = Ay + By ,

Cz = Az + Bz .

Scalar product
Since i i = j j = k k = 1 and that i j = j k = k i = 0, the scalar product of two vectors can be
written as,
A B = (Ax i + Ay j + Az k) (Bx i + By j + Bz k) = Ax Bx + Ay By + Az Bz .
Note that, A A = A2 = A2x + A2y + A2z , which is consistent with Pythagoras theorem.
Vector product
Here, i i = j j = k k = 0 and i j = k, j k = i, and k i = j. Thus,
AB

(Ax i + Ay j + Az k) (Bx i + By j + Bz k)

(Ay Bz Az By )i + (Az Bx Ax Bz )j + (Ax By Ay Bx )k = Ax

Bx

Triple product
The triple product A (B C) can be expressed as the following determinant

Ax Ay Az

A (B C) = Bx By Bz ,

Cx Cy Cz

j
Ay
By

Az .

Bz

which clearly is equal to zero whenever the vectors are linearly dependent (if the three vectors are linearly
dependent they must be co-planar and therefore the parallelepiped dened by the three vectors has zero
volume).

Vector Transformations
In many problems we will need to use dierent coordinate systems in order to describe dierent vector
quantities. The above operations, written in component form, only make sense once all the vectors involved
are described with respect to the same frame. In this section, we will see how the components of a vector
are transformed when we change the reference frame.
Consider two dierent orthogonal, right-hand sided, reference frames x1 , x2 , x3 and X1 , X2 , X3 . A vector A
in coordinate system x can be transformed to coordinate system X by considering the 9 angles that dene
the relationships between the two systems. (Only three of these angles are independent, a point we shall
return to later.)
Referring to a) in the gure we see the vector A, the x and X coordinate systems, the unit vectors i1 , i2 , i3 of
the x system and the unit vectors i1 , i2 , i3 of the X system; a) focuses on the transformation of coordinates
from x to X while b) focuses on the reverse transformation from X to x.

In the x coordinate system, the vector A, can be written as


A = A1 i1 + A2 i2 + A3 i3 ,

(3)

A = A1 i1 + A2 i2 + A3 i3 .

(4)

or, when referred to the frame X, as

Since the vector A remains the same regardless of our coordinate transformation
A = A1 i1 + A2 i2 + A3 i3 = A1 i1 + A2 i2 + A3 i3 ,

(5)

We can nd the components of the vector A in the transformed system in term of the components of A in
the original system by simply taking the dot product of this equation with the desired unit vector ij in the
X system so that
Aj = A1 ij i1 + A2 ij i2 + A3 ij i3

(6)

where Aj is the jth component of A in the X system. Repeating this operation for each component of A
in the X system results in the matrix form for A


A1
i i i i

1 1 1 2

A2 = i2 i1 i2 i2


A3
i3 i1 i3 i2

i1 i3

A1

i2 i3 A2 .

i3 i3
A3

The above expression is the relationship that expresses how the components of a vector in one coordinate
system relate to the components of the same vector in a dierent coordinate system.
Referring to the gure, we see that ij ii is equal to the cosine of the angle between ij and ii which is j i ; in
particular we see that i2 i1 = cos21 while i1 i2 = cos12 ; these angles are in general not equal. Therefore,
the components of the vector A are transformed from the x coordinate system to the X system through
the transformation
Aj = A1 cos j1 + A2 cos j2 + A3 cos j3

(7)

where the coecients relating the components of A in the two coordinate systems are the various direction
cosines of the angles between the coordinate directions.
The above relations for the transformation of A from the x to the X system can be written in matrix form
as

A1

cos (11 )


A = A2 = cos (21 )


A3
cos (31 )

cos (12 )
cos (22 )
cos (32 )

cos (13 )

A1

cos (23 ) A2 .

cos (33 )
A3

(8)

We use the symbol A to denote the components of the vector A in the system. Of course the vector A is
unchanged by the transformation. We introduce the symbol [T ] for the transformation matrix from x to X.
10

This relationship, which expresses how the components of a vector in one coordinate system relate to the
components of the same vector in a dierent coordinate system, is then written

A = [T ]A.

(9)

where [T ] is the transformation matrix.

We now consider the process that transforms the vector A from the X system to the x system.

A1

cos (11 )

cos (12 )


A = A2 = cos (21 )


A3
cos (31 )

cos (22 )
cos (32 )

cos (13 )

A1

cos (23 ) A2 .

cos (33 )
A3

(10)

By comparing the two coordinate transformations shown in a) and b), we see that cos(12 )=cos(21 ), and
that therefore the matrix element of magnitude cos(12 ) which appears in the 12 position in the transfor
mation matrix from x to X now appears in the 21 position in the matrix which transforms A from X to
x. This pattern is repeated for all o-diagonal elements. The diagonal elements remain unchanged since
cos(ii )=cos(ii ). Thus the matrix which transposes the vector A in the Xsystem back to the x system is
the transpose of the original transformation matrix,

A = [T ]T A.

(11)

where [T ]T is the transpose of [T ]. (A transpose matrix has the rows and columns reversed.)
Since transforming A from x to X and back to x results in no change, the matrix [T ]T is also [T ]1 the
inverse of [T ] since
A = [T ]T [T ]A = [T ]1 [T ]A = [I]A = A.

(12)

where [I] is the identity matrix

I= 0

0
1
0

(13)

This is a remarkable and useful property of the transformation matrix, which is not true in general for any
matrix.
Example

Coordinate transformation in two dimensions

Here, we apply for illustration purposes, the above expressions to a two-dimensional example. Consider the
change of coordinates between two reference frames xy, and x y , as shown in the diagram.

11

The angle between i and i is . Therefore, i i = cos . Similarly, j i = cos(/2 ) = sin ,


i j = cos(/2 + ) = sin , and j j = cos . Finally, the transformation matrix [T ] is

cos (11 ) cos (12 )


cos
sin
=
,
[T ] =
cos (21 ) cos (22 )
sin cos
and we can write,

A1

A2

= [T ]

A1
A2

and

A1
A2

= [T ]T

A1
A2

Therefore,
A1 = A1 cos + A2 sin

(14)

A2 = A1 sin + A2 cos .

(15)

For instance, we can easily check that when = /2, the above expressions give A1 = A2 , and A2 = A1 ,

as expected.

An additional observation can be made. If in three dimensions, we rotate the x, y, z coordinate system about

the z axis, as shown in a) leaving the z component unchanged,

12

the transformation matrix becomes

cos (11 ) cos (12 ) 0

[T ] = cos (21 ) cos (22 ) 0

0
0
1

cos



= sin

0

sin
cos
0

0 .

Analogous results can be obtained for rotation about the x axis or rotation about the y axis as shown in b)
and c).

Sequential Transformations; Euler Angles


The general orientation of a coordinate system can be described by a sequence of rotations about coordinate
axis. One particular set of such rotations leads to a description particularly convenient for describing the
motion of a three-dimensional rigid body in general spinning motion, call Euler angles. We shall treat
this topic in Lecture 28. For now, we examine how this rotation ts into our general study of coordinate
transformations. A coordinate description in terms of Euler angles is obtained by the sequential rotation of
axis as shown in the gure; the order of transformation makes a dierence.
To develop the description of this motion, we use a series of transformations of coordinates. The nal result is
shown below. This is the coordinate system used for the description of motion of a general three-dimensional
rigid body such as a top described in body-xed axis. To identify the new position of the coordinate axes
as a result of angular displacement through the three Euler angles, we go through a series of coordinate
rotations.

13

First, we rotate from an initial X, Y, Z system into an x , y , z system through a rotation about the Z, z
axis.

cos



y = sin


z
0

sin 0
cos
0

0 Y = [T1 ] Y .

1
Z
Z

The resulting x , y coordinates remain in the X, Y plane. Then, we rotate about the x axis into the
x , y , z system through an angle . The x axis remains coincident with the x axis. The axis of rotation
for this transformation is called the line of nodes. The plane containing the x , y coordinate is now tipped
through an angle relative to the original X, Y plane. coordinates

x
1
0
0
x
x

y = 0 cos sin y = [T2 ] y .

z
0 sin cos
z
z
And nally, we rotate about the z , z system through an angle into the x, y, z system. The z axis is
called the spin axis. It is coincident with the z axis.

x
cos sin 0
x

y = sin cos 0 y

z
0
0
1
z

= [T3 ] y

The nal coordinate system used to describe the position of the body is shown below. The angle is called

the spin; the angle is called the precession; the angle is called the nutation. The total transformation is

given by

x
X

y = [T3 ][T2 ][T1 ] Y .

z
Z

14

Euler angles are not always dened in exactly this manner, either the notation or the order of rotations can
dier. The particular transformation used in any example should be clearly described.

References
[1] J.B. Marion and S.T. Thornton, Classical Dynamics of Particles and Systems, Harcourt Brace, 1995.
[2] D. Kleppner and R.J. Kolenkow, An Introduction to Mechnics, McGraw Hill, 1973.

15

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

S. Widnall, J. Peraire
16.07 Dynamics
Fall 2009
Version 2.0

Lecture L4 - Curvilinear Motion. Cartesian Coordinates


We will start by studying the motion of a particle. We think of a particle as a body which has mass,
but has negligible dimensions. Treating bodies as particles is, of course, an idealization which involves an
approximation. This approximation may be perfectly acceptable in some situations and not adequate in
some other cases. For instance, if we want to study the motion of planets, it is common to consider each
planet as a particle. This simplication is not adequate if we wish to study the precession of a gyroscope or
a spinning top.

Kinematics of curvilinear motion


In dynamics we study the motion and the forces that cause, or are generated as a result of, the motion.
Before we can explore these connections we will look rst at the description of motion irrespective of the
forces that produce them. This is the domain of kinematics. On the other hand, the connection between
forces and motions is the domain of kinetics and will be the subject of the next lecture.

Position vector and Path


We consider the general situation of a particle moving in a three dimensional space. To locate the position of
a particle in space we need to set up an origin point, O, whose location is known. The position of a particle
A, at time t, can then be described in terms of the position vector, r, joining points O and A. In general,
this particle will not be still, but its position will change in time. Thus, the position vector will be a function
of time, i.e. r(t). The curve in space described by the particle is called the path, or trajectory.

We introduce the path or arc length coordinate, s, which measures the distance traveled by the particle along
the curved path. Note that for the particular case of rectilinear motion (considered in the review notes) the
arc length coordinate and the coordinate, s, are the same.

Using the path coordinate we can obtain an alternative representation of the motion of the particle. Consider
that we know r as a function of s, i.e. r(s), and that, in addition we know the value of the path coordinate
as a function of time t, i.e. s(t). We can then calculate the speed at which the particle moves on the path
simply as v = s ds/dt. We also compute the rate of change of speed as at = s = d2 s/dt2 .
We consider below some motion examples in which the position vector is referred to a xed cartesian
coordinate system.
Example

Motion along a straight line in 2D

Consider for illustration purposes two particles that move along a line dened by a point P and a unit vector
m. We further assume that at t = 0, both particles are at point P . The position vector of the rst particle is
given by r 1 (t) = r P + mt = (rP x + mx t)i + (rP y + my t)j, whereas the position vector of the second particle
is given by r 2 (t) = r P + mt2 = (rP x + mx t2 )i + (rP y + my t2 )j.

Clearly the path for these two particles is the same, but the speed at which each particle moves along the
path is dierent. This is seen clearly if we parameterize the path with the path coordinate, s. That is,
we write r(s) = r P + ms = (rP x + mx s)i + (rP y + my s)j. It is straightforward to verify that s is indeed
the path coordinate i.e. the distance between two points r(s) and r(s + s) is equal to s. The two
motions introduced earlier simply correspond to two particles moving according to s1 (t) = t and s2 (t) = t2 ,
respectively. Thus, r 1 (t) = r(s1 (t)) and r 2 (t) = r(s2 (t)).

It turns out that, in many situations, we will not have an expression for the path as a function of s. It is
in fact possible to obtain the speed directly from r(t) without the need for an arc length parametrization of
the trajectory.

Velocity Vector
We consider the positions of the particle at two dierent times t and t + t, where t is a small increment
of time. Let r = r(r + t) r(t), be the displacement vector as shown in the diagram.

The average velocity of the particle over this small increment of time is
v ave =

r
,
t

which is a vector whose direction is that of r and whose magnitude is the length of r divided by t. If
t is small, then r will become tangent to the path, and the modulus of r will be equal to the distance
the particle has moved on the curve s.
The instantaneous velocity vector is given by
v = lim

t0

r
dr(t)

r ,
t
dt

(1)

and is always tangent to the path. The magnitude, or speed, is given by


v = |v | = lim

t0

s
ds

s .
t
dt

Acceleration Vector
In an analogous manner, we can dene the acceleration vector. Particle A at time t, occupies position
r(t), and has a velocity v(t), and, at time t + t, it has position r(t + t) = r(t) + r, and velocity
v(t + t) = v(t) + v. Considering an innitesimal time increment, we dene the acceleration vector as the
derivative of the velocity vector with respect to time,
v
dv
d2 r

= 2 .
t0 t
dt
dt

a = lim

(2)

We note that the acceleration vector will reect the changes of velocity in both magnitude and direction.
The acceleration vector will, in general, not be tangent to the trajectory (in fact it is only tangent when the
velocity vector does not change direction). A sometimes useful way to visualize the acceleration vector is to
3

translate the velocity vectors, at dierent times, such that they all have a common origin, say, O . Then,
the heads of the velocity vector will change in time and describe a curve in space called the hodograph. We
then see that the acceleration vector is, in fact, tangent to the hodograph at every point.

Expressions (1) and (2) introduce the concept of derivative of a vector. Because a vector has both magnitude
and direction, the derivative will be non-zero when either of them changes (see the review notes on
vectors). In general, the derivative of a vector will have a component which is parallel to the vector itself,
and is due to the magnitude change; and a component which is orthogonal to it, and is due to the direction
change.
Note

Unit tangent and arc-length parametrization

The unit tangent vector to the curve can be simply calculated as


et = v/v.
It is clear that the tangent vector depends solely on the geometry of the trajectory and not on the speed
at which the particle moves along the trajectory. That is, the geometry of the trajectory determines the
tangent vector, and hence the direction of the velocity vector. How fast the particle moves along the
trajectory determines the magnitude of the velocity vector. This is clearly seen if we consider the arc-length
parametrization of the trajectory r(s). Then, applying the chain rule for dierentiation, we have that,
v=

dr
dr ds
=
= et v ,
dt
ds dt

where, s = v, and we observe that dr/ds = et . The fact that the modulus of dr/ds is always unity indicates
that the distance traveled, along the path, by r(s), (recall that this distance is measured by the coordinate
s), per unit of s is, in fact, unity!. This is not surprising since by denition the distance between two
neighboring points is ds, i.e. |dr| = ds.

Cartesian Coordinates
When working with xed cartesian coordinates, vector dierentiation takes a particularly simple form. Since
the vectors i, j, and k do not change, the derivative of a vector A(t) = Ax (t)i + Ay (t)j + Az (t)k, is simply
(t) = Ax (t)i + Ay (t)j + Az (t)k. That is, the components of the derivative vector are simply the derivatives
A
of the components.
4

Thus, if we refer the position, velocity, and acceleration vectors to a xed cartesian coordinate system, we
have,
r(t)

= x(t)i + y(t)j + z(t)k

(3)

v(t)

+ z(t)k

= r (t)
= vx (t)i + vy (t)j + vz (t)k = x (t)i + y(t)j

(4)

a(t)

ax (t)i + ay (t)j + az (t)k = v x (t)i + v y (t)j + v z (t)k = v (t)

Here, the speed is given by v =

(5)

vx2 + vy2 + vz2 , and the magnitude of the acceleration is a = a2x + a2y + a2z .

The advantages of cartesian coordinate systems is that they are simple to use, and that if a is constant, or
a function of time only, we can integrate each component of the acceleration and velocity independently as
shown in the ballistic motion example.
Example

Circular Motion

We consider motion of a particle along a circle of radius R at a constant speed v0 . The parametrization of
a circle in terms of the arc length is
r(s) = R cos(

s
s
)i + R sin( )j .
R
R

Since we have a constant speed v0 , we have s = v0 t. Thus,


r(t) = R cos(

v0 t
v0 t
)i + R sin(
)j .
R
R

The velocity is
v(t) =

dr(t)
v0 t
v0 t
= v0 sin(
)i + v0 cos(
)j ,
dt
R
R
5

which, clearly, has a constant magnitude |v| = v0 . The acceleration is,


dr(t)
v2
v0 t
v2
v0 t
= 0 cos(
)i 0 sin(
)j .
dt
R
R
R
R

a(t) =

Note that, the acceleration is perpendicular to the path (in this case it is parallel to r), since the velocity
vector changes direction, but not magnitude.
We can also verify that, from r(s), the unit tangent vector, et , could be computed directly as
et =

dr(s)
s
s
v0 t
v0 t
= sin( )i + cos( ) = sin(
)i + cos(
)j .
ds
R
R
R
R

Example

Motion along a helix

The equation r(t) = R cos ti + R sin tj + htk, denes the motion of a particle moving on a helix of radius R,
and pitch 2h, at a constant speed. The velocity vector is given by
v=

dr
= R sin ti + R cos tj + hk ,
dt

and the acceleration vector is given by,


a=

dv
= R cos ti + R sin tj .
dt

In order to determine the speed at which the particle moves we simply compute the modulus of the velocity
vector,
v = |v| =

R2 sin2 t + R2 cos2 t + h2 =

R 2 + h2 .

If we want to obtain the equation of the path in terms of the arc-length coordinate we simply write,

ds = |dr| = vdt = R2 + h2 dt .
Integrating, we obtain s = s0 +

R2 + h2 t, where s0 corresponds to the path coordinate of the particle

at time zero. Substituting t in terms of s, we obtain the expression for the position vector in terms of the

arc-length coordinate. In this case, r(s) = R cos(s/ R2 + h2 )i + R sin(s/ R2 + h2 )j + hs/ R2 + h2 k. The


gure below shows the particle trajectory for R = 1 and h = 0.1.

2
1.5
1
0.5
0
1
1

0.5

0.5

0
0.5

0.5

1 1

Example

Ballistic Motion

Consider the free-ight motion of a projectile which is initially launched with a velocity v 0 = v0 cos i +
v0 sin j. If we neglect air resistance, the only force on the projectile is the weight, which causes the projectile
to have a constant acceleration a = gj. In component form this equation can be written as dvx /dt = 0
and dvy /dt = g. Integrating and imposing initial conditions, we get
vx = v0 cos ,

vy = v0 sin gt ,

where we note that the horizontal velocity is constant. A further integration yields the trajectory
x = x0 + (v0 cos ) t,

1
y = y0 + (v0 sin ) t gt2 ,
2

which we recognize as the equation of a parabola.

The maximum height, ymh , occurs when vy (tmh ) = 0, which gives tmh = (v0 /g) sin , or,
ymh = y0 +

v02 sin2
.
2g

The range, xr , can be obtained by setting y = y0 , which gives tr = (2v0 /g) sin , or,
xr = x0 +

2v02 sin cos


v 2 sin(2)
= x0 + 0
.
g
g

We see that if we want to maximize the range xr , for a given velocity v0 , then sin(2) = 1, or = 45o .
Finally, we note that if we want to model a more realistic situation and include aerodynamic drag forces of
the form, say, v 2 , then we would not be able to solve for x and y independently, and this would make the
problem considerably more complicated (usually requiring numerical integration).

ADDITIONAL READING
J.L. Meriam and L.G. Kraige, Engineering Mechanics, DYNAMICS, 5th Edition
2/1, 2/3, 2/4

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

S. Widnall, J. Peraire
16.07 Dynamics
Fall 2008
Version 2.0

Lecture L5 - Other Coordinate Systems


In this lecture, we will look at some other common systems of coordinates. We will present polar coordinates
in two dimensions and cylindrical and spherical coordinates in three dimensions. We shall see that these
systems are particularly useful for certain classes of problems.

Polar Coordinates (r )
In polar coordinates, the position of a particle A, is determined by the value of the radial distance to the
origin, r, and the angle that the radial line makes with an arbitrary xed line, such as the x axis. Thus, the
trajectory of a particle will be determined if we know r and as a function of t, i.e. r(t), (t). The directions
of increasing r and are dened by the orthogonal unit vectors er and e .

The position vector of a particle has a magnitude equal to the radial distance, and a direction determined
by er . Thus,
r = rer .

(1)

Since the vectors er and e are clearly dierent from point to point, their variation will have to be considered

when calculating the velocity and acceleration.

Over an innitesimal interval of time dt, the coordinates of point A will change from (r, ), to (r + dr, + d)

as shown in the diagram.

We note that the vectors er and e do not change when the coordinate r changes. Thus, der /dr = 0 and
de /dr = 0. On the other hand, when changes to + d, the vectors er and e are rotated by an angle
d. From the diagram, we see that der = de , and that de = der . This is because their magnitudes in
the limit are equal to the unit vector as radius times d in radians. Dividing through by d, we have,
der
= e ,
d

and

de
= er .
d

Multiplying these expressions by d/dt , we obtain,


der d
der

= e ,
d dt
dt

Note

and

de
= er .
dt

(2)

Alternative calculation of the unit vector derivatives

An alternative, more mathematical, approach to obtaining the derivatives of the unit vectors is to express
er and e in terms of their cartesian components along i and j. We have that
er

= cos i + sin j

= sin i + cos j .

Therefore, when we dierentiate we obtain,


der
= 0,
dr
de
= 0,
dr

der
d

de

= sin i + cos j e
= cos i sin j er .

Velocity vector
We can now derive expression (1) with respect to time and write
v = r = r er + r e r ,
or, using expression (2), we have
v = r er + r e .

(3)

Here, vr = r is the radial velocity component, and v = r is the circumferential velocity component. We

also have that v = vr2 + v2 . The radial component is the rate at which r changes magnitude, or stretches,
and the circumferential component, is the rate at which r changes direction, or swings.

Acceleration vector
Dierentiating again with respect to time, we obtain the acceleration
a = v = r er + r e r + r
e + r e + r e
Using the expressions (2), we obtain,
a = (
r r2 ) er + (r + 2r
) e ,

(4)

r r2 ) is the radial acceleration component, and a = (r + 2r


) is the circumferential
where ar = (

acceleration component. Also, we have that a = a2r + a2 .

Change of basis
In many practical situations, it will be necessary to transform the vectors expressed in polar coordinates to
cartesian coordinates and vice versa.

Since we are dealing with free vectors, we can translate the polar reference frame for a given point (r, ), to
the origin, and apply a standard change of basis procedure. This will give, for a generic vector A,

Ar
cos sin
Ax
cos sin
Ax
Ar

=
.

and
A
sin cos
Ay
Ay
sin
cos
A

Example

Circular motion

Consider as an illustration, the motion of a particle in a circular trajectory having angular velocity = ,
and angular acceleration = .

In polar coordinates, the equation of the trajectory is


1
= t + t2 .
2

r = R = constant,
The velocity components are
vr = r = 0,

v = r = R( + t) = v ,

and

and the acceleration components are,


2

v
ar = r r2 = R( + t)2 = ,
R

and

a = r + 2r
= R = at ,

where we clearly see that, ar an , and that a at .


In cartesian coordinates, we have for the trajectory,
1
x = R cos(t + t2 ),
2

1
y = R sin(t + t2 ) .
2

For the velocity,


1
vx = R( + t) sin(t + t2 ),
2

1
vy = R( + t) cos(t + t2 ) ,
2

and, for the acceleration,


1
1
ax = R(+t)2 cos(t+ t2 )R sin(t+ t2 ),
2
2

1
1
ay = R(+t)2 sin(t+ t2 )+R cos(t+ t2 ) .
2
2

We observe that, for this problem, the result is much simpler when expressed in polar (or intrinsic) coordi
nates.

Example

Motion on a straight line

Here we consider the problem of a particle moving with constant velocity v0 , along a horizontal line y = y0 .

Assuming that at t = 0 the particle is at x = 0, the trajectory and velocity components in cartesian
coordinates are simply,
x = v0 t

y = y0

vx = v0

vy = 0

ax = 0

ay = 0 .
4

In polar coordinates, we have,

vr = r = v0 cos

y0
)
v0 t
v = r = v0 sin

ar = r r2 = 0

a = r + 2r
=0 .

r=

v02 t2 + y02

= tan1 (

Here, we see that the expressions obtained in cartesian coordinates are simpler than those obtained using
polar coordinates. It is also reassuring that the acceleration in both the r and direction, calculated from
the general two-term expression in polar coordinates, works out to be zero as it must for constant velocitystraight line motion.

Example

Spiral motion (Kelppner/Kolenkow)

A particle moves with = = constant and r = r0 et , where r0 and are constants.

We shall show that for certain values of , the particle moves with ar = 0.
a

= (
r r2 )er + (r + 2r
)e
=

( 2 r0 et r0 et 2 )er + 2r0 et e

If = , the radial part of a vanishes. It seems quite surprising that when r = r0 et , the particle moves
with zero radial acceleration. The error is in thinking that r makes the only contribution to ar ; the term
r2 is also part of the radial acceleration, and cannot be neglected.
The paradox is that even though ar = 0, the radial velocity vr = r = r0 et is increasing rapidly in time.
In polar coordinates

vr =

ar (t)dt ,

because this integral does not take into account the fact that er and e are functions of time.

Equations of Motion
In two dimensional polar r coordinates, the force and acceleration vectors are F = Fr er + F e and
a = ar er + a e . Thus, in component form, we have,
Fr

= m ar = m (
r r2 )

= m a = m (r + 2r
) .

Cylindrical Coordinates (r z)
Polar coordinates can be extended to three dimensions in a very straightforward manner. We simply add
the z coordinate, which is then treated in a cartesian like manner. Every point in space is determined by
the r and coordinates of its projection in the xy plane, and its z coordinate.

The unit vectors er , e and k, expressed in cartesian coordinates, are,


er

= cos i + sin j

= sin i + cos j

and their derivatives,


e r = e ,

e = er ,

k = 0 .

The kinematic vectors can now be expressed relative to the unit vectors er , e and k. Thus, the position
vector is
r = r er + z k ,
and the velocity,
v = r er + r e + z k ,
where vr = r, v = r, vz = z, and v =

vr2 + v2 + vz2 . Finally, the acceleration becomes

a = (
r r2 ) er + (r + 2r
) e + zk ,
where ar = r r2 , a = r + 2r
, az = z, and a =

a2r + a2 + az2 .
6

Note that when using cylindrical coordinates, r is not the modulus of r. This is somewhat confusing, but it
is consistent with the notation used by most books. Whenever we use cylindrical coordinates, we will write

|r| explicitly, to indicate the modulus of r, i.e. |r| = r2 + z 2 .

Equations of Motion
In cylindrical rz coordinates, the force and acceleration vectors are F = Fr er + F e + Fz ez and a =
ar er + a e + az ez . Thus, in component form we have,
Fr

= m ar = m (
r r2 )

= m a = m (r + 2r
)

Fz

= m az = m z .

Spherical Coordinates (r )
In spherical coordinates, we utilize two angles and a distance to specify the position of a particle, as in the
case of radar measurements, for example.

The unit vectors written in cartesian coordinates are,


er

= cos cos i + sin cos j + sin k

= sin i + cos j

= cos sin i sin sin j + cos k

The derivation of expressions for the velocity and acceleration follow easily once the derivatives of the unit
vectors are known. In three dimensions, the geometry is somewhat more involved, but the ideas are the
same. Here, we give the results for the derivatives of the unit vectors,
e r = cos e + e ,

e = cos er + sin e ,

e = er sin e ,

and for the kinematic vectors


r

= r er

= rer + r cos e + r e

= (
r r2 cos2 r 2 ) er
+ (2r
cos + r cos 2r sin ) e
+ (2r
+ r 2 sin cos + r) e .

Equations of Motion
Finally, in spherical r coordinates, we write F = Fr er + F e + F e and a = ar er + a e + a e . Thus,
Fr

= m ar = m (
r r2 cos2 r 2 )

= m a = m (2r
cos + r cos 2r sin )

= m a = m (2r
+ r 2 sin cos + r) .

Application Examples
We will look at some applications of Newtons second law, expressed in the dierent coordinate systems that
have been introduced. Recall that Newtons second law
F = ma ,

(5)

is a vector equation which is valid for inertial observers.


In general, we will be interested in determining the motion of a particle given that we know the external
forces. Equation (5), written in terms of either velocity or position, is a dierential equation. In order
to calculate the velocity and position as a function of time we will need to integrate this equation either
analytically or numerically. On the other hand, the reverse problem of computing the forces given motion is
much easier and only requires direct evaluation of (5). Is is also common to have mixed type problems, in
which we know some components of the force and some components of the acceleration. The goal is then to
determine the remaining unknown terms.
While no general rules can be given regarding the appropriate choice of a coordinate system, we note
that intrinsic coordinates are particularly useful in constrained problems, where the trajectory is known
beforehand.
Example

Aircraft ying on a helix

A 10, 000 lb aircraft is descending on a cylindrical helix. The rate of descent is z = 10ft/s, the speed is
v = 211 ft/s, and = 3o 0.05rad/s. This is standard for gas turbine powered aircraft. We want to know
the force on the aircraft and the radius of curvature of the path.
8

We have,
v = rer + re + ze
z = vet
Since, r = R, r = 0. Therefore, 211 =

(0.05R)2 + 102 , or R = 4, 215 ft. For the acceleration,

v2
a = (
r r2 )er + (r + 2r
)e + zez = vet + en ,

and, considering only the non-zero terms,


v2
a = R2 er = en .

We see that en = er , and that,


2

a = (0.05)2 4, 215 = 10.54 ft/s =

v2
,

211
= 4, 225 ft .
10.54

The normal force on the aircraft is


Fn = man =

10, 000
10.54 = 3, 273 lb ,
32

and nally, the lift, L, is


L = 3, 273 er + 10, 000 ez lb .

Here we see that r which means that the helix is very tight.

The angle of descent is calculated as sin = z/v,

or, = 2.72o . This angle is sometimes called the


pitch of the helix.

Example

Pendulum

Now, we consider a simple pendulum consisting of a mass, m, suspended from a string of length l and
negligible mass.

We can formulate the problem in polar coordinates, and noting that r = l (constant), write for the r and
components,
mg cos T
mg sin

= ml2
=

ml ,

(6)

where T is the tension on the string. If we restrict the motion to small oscillations, we can approximate
sin , and the -equation becomes
g
+ = 0 .
l
Integrating we obtain the general solution,

g
g
(t) = C1 cos(
t) + C2 sin(
t) ,
l
l
where the constants C1 and C2 are determined by the initial conditions. Thus, if (0) = max ,

g
(t) = max cos(
t) .
l
10

Example

Aircraft ying a perfect loop (Hollister)

Consider an aircraft ying a perfect loop, i.e. a circle in the vertical plane. Assume that the engine thrust
exactly cancels the aerodynamic drag so that the lift and gravity are the only unbalanced forces on the
aircraft. This assumption makes the problem into the same dynamical model that we have used in the
previous example.

Since the lift, L, is perpendicular to the ight path, we have that the force on the aircraft, in normal and
tangential components, is
F = mg sin et + (L mg cos ) en .
Thus,
at
an

= v = r = g sin
v2
L
=
g cos .
=
R
m

(7)

Since, v dv = at ds = at R d = Rg sin d. Thus, integrating,


v 2 = v02 + 2Rg(cos 1) ,

(8)

where v0 is the velocity at the bottom of the loop when = 0. To be able to go over the top we need v > 0

when = . This means that we need v0 > 2 Rg.

Note that for v0 < 2 Rg, we can calculate the maximum angle the aircraft can reach, max . If we set v = 0
when = max , we have,

v02
).
2Rg
The necessary lift, L, can be calculated as a function of . From (7) and (8), we have
max = cos1 (1

L
v2
v2
=
+ g cos = 0 + 3g cos 2g .
m
R
R
We have that, in order for to go from 0 to , the aircraft has to have a range of lift capability that extends

over 5g.

It turns out that most aircraft do not have this capability and consequently do not y perfect loops.

11

ADDITIONAL READING
J.L. Meriam and L.G. Kraige, Engineering Mechanics, DYNAMICS, 5th Edition
2/6, 2/7, 3/5

12

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

S. Widnall, J. Peraire
16.07 Dynamics
Fall 2009
Version 2.0

Lecture L6 - Intrinsic Coordinates


In lecture L4, we introduced the position, velocity and acceleration vectors and referred them to a xed
cartesian coordinate system . Then we showed how they could be expressed in polar coordinates. While it
is clear that the choice of coordinate system does not aect the nal answer, we shall see that, in practical
problems, the choice of a specic system may simplify the calculations and/or improve the understanding
considerably. In many problems, it is useful to use a coordinate system that aligns with both the velocity
vector of the particlewhich is of course tangent to the particles trajectory and the normal to this
trajectory, forming a pair of orthogonal unit vectors. The unit vectors aligned with these two directions also
dene a third direction, call the binormal which is normal to both the velocity vector and the normal vector.
This can be obtained from these two unit vectors by taking their cross product. Such a coordinate system
is called an Intrinsic Coordinate System.

Intrinsic coordinates: Tangential, Normal and Binormal compo


nents.
We follow the motion of a point using a vector r(t) whose position along a known curved path is given by
the scalar function s(t) where s(t) is the arc length along the curve. We obtain the velocity v from the
time rate of change of the vector r(t) following the particle

v=

dr
dr ds
dr
=
= s
dt
ds dt
ds

We identify the scalar s as the magnitude of the velocity v and

(1)
dr
ds

as the unit vector tangent to the curve

at the point s(t). Therefore we have


v = vet ,

(2)

where r(t) is the position vector, v = s is the speed, et is the unit tangent vector to the trajectory, and s is
the path coordinate along the trajectory.

The unit tangent vector can be written as,


et =

dr
.
ds

The acceleration vector is the derivative of the velocity vector with respect to time. Since

(3)
dr
ds

depends only

on s, using the chain rule we can write

a=

dv
dr
d
= s + s
dt
ds
dt

dr
ds

= s

dr
d2 r
dr
d2 r
+ s 2 2 = v
+ v2 2 .
ds
ds
ds
ds

(4)

The second derivative d2 r/ds2 is another property of the path. We shall see that it is relate to the radius of
curvature.
Taking the time derivative of Equation (2), an alternate expression can be written in terms of the unit vector
et as
a=

dv
det
et + v
.
dt
dt

(5)

The vector et is the local unit tangent vector to the curve which changes from point to point. Consequently,

the time derivative of et will, in general, be nonzero.

The time derivative of et can be written as,

det
det ds
det
=
=
v.
dt
ds dt
ds

(6)

In order to calculate the derivative of et , we note that, since the magnitude of et is constant and equal to
one, the only changes that et can have are due to rotation, or swinging.

When we move from s to s + ds, the tangent vector changes from et to et + det . The change in direction
can be related to the angle d.
2

The direction of det , which is perpendicular to et , is called the normal direction. On the other hand, the
magnitude of det will be equal to the length of et (which is one), times d. Thus, if en is a unit normal
vector in the direction of det , we can write
det = den .

(7)

det
d
1
=
en = en = en .
ds
ds

(8)

Dividing by ds yields,

Here, = d/ds is a local property of the curve, called the curvature, and = 1/ is called the radius of

curvature.

Note that in the picture, the sizes of det , ds, and d are exaggerated for illustration purposes and actually

represent the changes in the limit as ds (and also dt) approach zero.

Note

Curvature and radius of curvature

We consider here two tangent vectors et and e + det , separated by a small ds and having an angle between
them of d. If we draw perpendiculars to these two vectors, they will intersect at a point, say, O . Because
the two lines meeting at O are perpendicular to each of the tangent vectors, the angle between them will
be the same as the angle between et and e + det , d. The point O is called the center of curvature, and the
distance, , between O and A is the radius of curvature. Thus, from the sketch, we have that ds = d, or
d/ds = = 1/.

Intuitively, we can see that each innitesimal arc, ds, can be represented by a circle segment of radius
having its center at the center of curvature. It is clear that both the radius of curvature and the center of
curvature are functions of s, and consequently they change from point to point.
There are two limiting cases which are of interest. When the trajectory is a circle, the center of curvature
does not change and coincides with the center of the circle, and the radius of curvature is equal to the radius

of the circle. On the other hand, when the trajectory is a straight line, the curvature is zero, and the radius
of curvature is innite. Note also, that, in this case, the derivative of et is always zero, and the normal
direction is not dened.

Going back to expression (6), we have that


det
d
v
=
v en = en = en .
dt
ds

(9)

Finally, we have that the acceleration, can be written as


a=

dv
v2
et + en = at et + an en .
dt

(10)

Here, at = v, is the tangential component of the acceleration, and an = v 2 /, is the normal component of
the acceleration. Since an is the component of the acceleration pointing towards the center of curvature, it is
sometimes referred to as centripetal acceleration. When at is nonzero, the velocity vector changes magnitude,
or stretches. When an is nonzero, the velocity vector changes direction, or swings. The modulus of the total

acceleration can be calculated as a = a2t + a2n .


Example
A ball is ejected horizontally from the tube with a speed v0 . The only acceleration on the ball is due to
gravity. We want to determine the radius of curvature of the trajectory just after the ball is released.

The simplest way to determine the radius of curvature is to note that, initially, the only nonzero component
of the acceleration will be in the normal direction, i.e. an = g. Thus, from an = v02 /, we have that,
=

v02
.
g
4

Alternatively, we can obtain an equation for the trajectory of the form y = f (x) and use expression (11) to
calculate the curvature. The trajectory is given as,
x = v0 t
1
y = gt2 .
2
Thus, eliminating t, we have
y=

g 2
x .
2v02

At x = 0, dy/dx = 0, d2 y/dx2 = g/v02 , and the above expression gives, = v02 /g, as expected.

Note

Relationship between s, v and at

The quantities s, v and at are related in the same manner as the quantities s, v and a for rectilinear motion.
In particular we have that v = s, at = v, and at ds = v dv. This means that if we have a way of knowing at ,
we may be able to integrate the tangential component of the motion independently. We will be exploiting
these relations in the future.

The vectors et and en , and their respective coordinates t and n, dene two orthogonal directions. The plane
dened by these two directions, is called the osculating plane. This plane changes from point to point, and
can be thought of as the plane that locally contains the trajectory (Note that the tangent is the current
direction of the velocity, and the normal is the direction into which the velocity is changing).
In order to dene a right-handed set of axes we need to introduce an additional unit vector which is orthogonal
to et and en . This vector is called the binormal, and is dened as eb = et en .

At any point in the trajectory, the position vector, the velocity and acceleration can be referred to these
axes. In particular, the velocity and acceleration take very simple forms,
v

= vet

a = ve
t+

v2
en .

The diculty of working with this reference frame stems from the fact that the orientation of the axis depends
on the trajectory itself. The position vector, r, needs to be found by integrating the relation dr/dt = v as
5

follows,

r = r0 +

v dt ,
0

where r 0 = r(0) is given by the initial condition.


We note that, by construction, the component of the acceleration along the binormal is always zero.
When the trajectory is planar, the binormal stays constant (orthogonal to the plane of motion). However,
when the trajectory is a space curve, the binormal changes with s. It can be shown (see note below) that the
derivative of the binormal is always along the direction of the normal. The rate of change of the binormal
with s is called the torsion, . Thus,
deb
= en
ds

or,

deb
= v en .
dt

We see that whenever the torsion is zero, the trajectory is planar, and whenever the curvature is zero, the
trajectory is linear.
Note

Calculation of the radius of curvature and torsion for a trajectory

In some situations the trajectory will be known as a curve of the form y = f (x). The radius of curvature in
this case can be computed according to the expression,
=

[1 + (dy/dx)2 ]3/2
.
|d2 y/dx2 |

(11)

This expression is not hard to derive. Try it! Since y = f (x) denes a planar curve, the torsion is zero.
On the other hand, if the trajectory is known in parametric form as a curve of the form r(t), where t can be
time, but also any other parameter, then the radius of curvature and the torsion can be computed as
(r r )3/2
,
(r r )(r r) (r r)2

=
where r = dr/dt, and r = d2 r/dt2 and

=
...
where r = d3 r/dt3 .

...
(r r ) r
|r r |2

(12)

Equations of Motion in Intrinsic Coordinates


Newtons second law is a vector equation, F = ma, which can now be written in intrinsic coordinates.

In tangent, normal and binormal components, tnb, we write F = Ft et + Fn en and a = at et + an en . We

observe that the positive direction of the normal coordinate is that pointing towards the center of curvature.

Thus, in component form, we have

Ft
Fn

= m at = m v = m s
v2
= m an = m

Note that, by denition, the component of the acceleration along the binormal direction, eb , is always zero,
and consequently the binormal component of the force must also be zero. This may seem surprising, at rst,
but recall that the tangent and normal directions are determined by the motion, and, hence, we can say that
the motion chooses the binormal direction to be always orthogonal to the applied force. In other words,
if we apply a force to a particle, the particle will experience an acceleration which is parallel to the force.
The normal direction is chosen so that the acceleration vector is always contained in the plane dened by
the tangent and the normal. Thus, the binormal is always orthogonal to the external force.
Intrinsic coordinates are sometimes useful when we are dealing with problems in which the motion is con
strained, such as a car on a roller coaster. The geometry of the trajectory is known, and, therefore, the
directions of the tangent, normal and binormal vectors are also known. In these cases it may be possible to
integrate the component of the equation of motion along the tangential direction (especially if there is no
friction), and then calculate, a-posteriori, the reaction force using the normal component of the equation of
motion.
Note (optional)

Frenet formulae

The Frenet formulae give us the variations of the unit vectors et , en and eb with respect to the path
coordinate s. The rst formula
det
1
= en ,
ds

has already been dened. Now, since eb is a unit vector, deb /ds will be orthogonal to eb . Hence, it will be
of the form,
deb
= bt et + bn en .
ds
If we perform the dot product of this expression with et , we obtain
bt =

deb
det
1
et =
eb = en eb = 0 .
ds
ds

The second equality follows from the fact that the derivative of et eb is zero, i.e. det eb + et deb = 0.
Therefore, only bn is nonzero. Dening bn = , we obtain the third Frenet formula
deb
= en .
ds
Finally, the second formula can be obtained in a similar manner (we leave the details as an exercise) and
gives,
den
1
= et + eb ,
ds

or, multiplying by v,
den
v
= et + veb .
dt

(13)

As we move along s, the osculating plane (and hence eb ) may rotate, making the curve non planar. As an
example, an aeroplane may be rolling as it ies along et . The derivative of eb is in the direction opposite to
7

en if the rotation is in the direction of a right hand screw, and this is taken as the positive direction for the
torsion.

Example

Simplied Aircraft Kinematics (W. M. Hollister)

The ight of an aircraft through the sky is an example of curvilinear motion. Think of et as the roll axis
aligned with the velocity vector of the aircraft. Think of eb as being the pitch axis. The lift is then directed
along en . The roll rate of the aircraft can be interpreted as v, and the pitch rate as v/. In order to
turn the aircraft out of the vertical plane, it is necessary to rotate the direction of the lift en so that there
is a component of acceleration out of the vertical plane. Neglecting gravity, the velocity vector along et
determines where the aircraft is going, and the lift along en determines where the velocity vector is going.
The roll rate determines how the lift vector will be rotated out of the osculating plane. As shown by equation
(13), the direction of the lift vector is changed by rolling v, as well as pitching v/.
Consider the following example. An aircraft follows a spiral path in the sky while doing a barrel roll. The
coordinates are given below, where v0 = 194 ft/s, = 0.4 rad/s, and h = 125 ft are constants.
x = v0 t
y

= h cos t

= h sin t

We have,

= v0 i h sin tj + h cos tk

=
v02 + h2 2 = 200ft/s v0

= h 2 cos tj h 2 sin tk

= h 2 = 20ft/s

Since v = 0, a = (v 2 /)en , or,


h 2 =

v2
,

v2
= 2000ft
h 2

,
v0
h
i
sin tj +
v
v
= cos tj sin tk
h
v0
=
i+
sin tj
v
v

et

en
eb

h
cos tk
v
v0
cos tk
v

Finally,
deb
v0
v0
=
cos tj +
sin tk
dt
v
v
which corresponds to a roll rate of
v =

v0
= 0.4rad/s
v

Example

Aircraft ying an imperfect loop (Hollister)

It is more common to y a loop keeping the normal acceleration, an , approximately constant at, say, ng

(n 3 4).

Let be the ight path angle which the velocity vector (or tangent) makes with the horizontal, and let be

the radius of curvature of the path. Then,

an
at

v2
= v = ng

= g sin .
=

From, v dv = at ds, with ds = d, we have,


v dv = at d = (g sin )(

v2
)d ,
ng

or, integrating,
n dv
= sin d ,
v

n ln v|v0 = cos |0 ,

v = v0 e n (1cos ) ,

and
=

v02 2 (1cos )
e n
.
ng

(14)

A sketch of this path is shown in the gure below.

PULL UP

RECOVERY

In going over the top of the imperfect loop, the aircraft does not go as high or loose as much velocity as
it does going over the perfect loop. Unlike the perfect loop case however, the aircraft does need to pull
up before, and recover after, the point of maximum altitude (see diagram). Note that the solution of this
example and the previous one would have been rather dicult using rectangular coordinates. Note also that
the form of the solution given by equation (14) is rather unusual, e.g. the radius of curvature is given as a
function of the attitude angle. A possible way to plot the trajectory starting from an initial position and
velocity would be to rst determine , and then draw a small circle segment with the appropriate . After
calculating the new position, the process would be repeated to draw the entire trajectory.

Example

Pursuit Problem

A classic problem in particle kinematics is the pursuit problem. This has an elegant history in mathematics
and practical applications in air combat. At its most basic, we have a target A traveling on a known
curve at constant speed. The target is pursued by B whose strategy is to turn such that her velocity vector
is always pointed at the target.

. We take the simplest version of this problem where the velocities of the two vehicles are equal: vA = vB
and the evading vehicle travels on a straight line in the positive x direction. Analyzing this problem using
intrinsic coordinates can give us some insight into the motion, into the formalism of tangential and normal
unit vectors and the role of radius of curvature of the path. We examine this problem at the instant where
the distance between A and B is given by R and the angle between their directions of travel is . (i.e. if the
particles are moving in the same direction, = 0.) We allow a time dt to elapse and examine the resulting
geometry. Of course to allow us to see the motion, we take dt to be rather large. Obviously we are talking
about a limiting process as dt > 0.
From the gure, we see that in a time dt, the vector R between the two points A and B changes direction
by an angle d = vA sin /Rdt. Therefore the direction of the tangential vector also changes by d. From
10

our earlier analysis the change of the tangential vector due to a change in ds is given by
det
d
1
= en
= en
ds
ds

(15)

Therefore, using the chain rule, the magnitude of radius of curvature at the point r(t) is given by
d
dt
ds
dt

1
d
=
=

ds

(16)

Now, ds/dt is the scalar velocity of the point B along the pursuit curve, while d/dt is the rate at which the
angle changes as a result of the motion of point A.
ds
dt
d
dt

= vB

vA sin
R

(17)
(18)

so that the radius of curvature at the point A is given by


1/ =

vA sin
vB R

(19)

2
Given this, the acceleration of particle B is easily found as a = vB
/.

ADDITIONAL READING
J.L. Meriam and L.G. Kraige, Engineering Mechanics, DYNAMICS, 5th Edition
2/5, 3/5 (normal and tangential coordinates only)

11

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

S. Widnall, J. Peraire
16.07 Dynamics
Fall 2009
Version 2.0

Lecture L7 - Relative Motion using Translating Axes


In the previous lectures we have described particle motion as it would be seen by an observer standing still
at a xed origin. This type of motion is called absolute motion. In many situations of practical interest,
we nd ourselves forced to describe the motion of bodies while we are simultaneously moving with respect
to a xed reference frame. There are many examples where such situations occur. The absolute motion of
a passenger inside an aircraft is best described if we rst consider the motion of the passenger relative to
the aircraft, and then the motion of the aircraft relative to the ground. If we try to track the motion of
aircraft in the airspace using satellites, it makes sense to rst consider the motion of the aircraft relative to
the satellite and then combine this motion with the motion of the satellite relative to the earths surface. In
this lecture we will introduce the ideas of relative motion analysis.

Types of observers
For the purpose of studying relative motion, we will consider four dierent types of observers (or reference
frames) depending on their motion with respect to a xed frame:
observers who do not accelerate or rotate, i.e. those who at most have constant velocity.
observers who accelerate but do not rotate
observers who rotate but do not translate
observers who accelerate and rotate
In this lecture we will consider the relative motion involving observers of the rst two types, and defer the
study of relative motion involving rotating frames to the next lecture.

Relative motion using translating axes


We consider two particles, A and B in curvilinear motions along two dierent paths. We describe their
motion with respect to a xed reference frame xyz with origin O and with unit vectors i, j and k, as before,
and call the motion relative to this frame absolute. The position of particle A is given by r A (t) and the
position of particle B is given by r B (t); both vectors are dened with respect to the xed reference frame
O. In addition, it is useful in many problems to ask how would B describe the motion of A and how would
this description be translate to the xed inertial coordinate system O?
1

In order to answer this question, we consider another translating reference frame attached to particle B,
x y z , with unit vectors i , j and k . Translating means that the angles between the axes xyz and x y z
do not change during the motion.

In the gure, we have chosen, for convenience, the axes xyz to be parallel to the axes x y z , but it should be
clear that one could have non-parallel translating axes. (By our ground rules, these axis must not rotate,
i.e the angles between them must not change; that possibility will be considered in a subsequent lecture.)
The position vector r A/B denes the position of A with respect to point B in the reference frame xyz.
The subscript notationA/B means A relative to B. The positions of A and B relative to the absolute
frame are given by the vectors r A and r B , respectively. Thus, we have
r A = r B + r A/B .
If we derive this expression with respect to time, we obtain
r A = r B + r A/B

or

v A = v B + v A/B ,

which relates the absolute velocities v A and v B to the relative velocity of A as observed by B. Dierentiating
again, we obtain an analogous expression for the accelerations,
r A = r B + r A/B

or

aA = aB + aA/B .

We will now reverse the roles of A and B , asking how would A describe the motion of B? We now attach
the reference frame x y z to A, and then we can observe B from A. The relative motion of B as seen by A
is now denoted r B/A , the position of B as seen by A.

The same arguments as before will give us,


r B = r A + r B/A ,

v B = v A + v B/A ,

aB = aA + aB/A .

Comparing these expressions with those above, we see that

r B/A = r A/B ,

v B/A = v A/B ,

aB/A = aA/B ,

as expected.
One important observation is that, whenever the moving system, say A, has a constant velocity relative
to the xed system, O, then the acceleration seen by the two observers is the same, i.e., if aA = 0, then
aB = aB/A . We shall see that this broadens the application of Newtons second law to systems which have
a constant absolute velocity.
Example

Glider in cross wind

Consider a glider ying above the edge of a cloud which is aligned North/South (360o ). The glider is ying
horizontally, and the cloud is stationary with respect to the ground. At the altitude of the glider ight, there
is a wind velocity, v w , of magnitude 52 knots coming from the direction 240o , as shown in the sketch.

North

Edge of Cloud
Cloud

The glider, on the other hand, is ying at a speed v G/w relative to the moving air mass and at an orientation
given by his compass heading.

In order to determine the speed of the glider with respect to the ground, we consider a reference frame
moving with the wind speed and write
v G = v w + v G/w ,
For these conditions, to balance the eect of wind speed and glider motion so that the glider stays at the
edge of the cloud, we require
v G/w Sin(360o heading) = v w Cos(30o )

(1)

Using this diagram, by adding and subtracting the vector components of the various velocities, we can ask
questions such as: if the glider is ying relative to the wind (v G/w )at a speed of 90 knots, what angle must
it make with the groundits absolute course headingto stay at the edge of the cloud? the glider ies with
a heading of 330o and a speed (relative to the wind) of 90 knots, it will stay at the edge of the cloud with a
ground speed of 104 = (26 + 78) knots north;
By inspection of the diagram below, we can also verify the following situations: if the glider ies with a
heading of 330o and a speed (relative to the wind) of 90 knots, it will stay at the edge of the cloud with a

ground speed of 104 = (26 + 78) knots north; if, on the other hand, the glider ies with a heading of 210o at
90 knots, the glider will stay at the edge of the cloud with a ground speed of 52 knots south; if the glider ies
into the wind at 240o , with a speed equal to 52 knots, it will remain stationary with respect to the ground;
nally, the lowest speed at which the glider can y to stay at the edge of the cloud is 45 knots and this will
be possible for a heading of 270o . The ground speed in this case will be 26 knots.

Example

Aircraft towing glider (Meriam)

Airplane A is ying horizontally with a constant speed of 200km/h, and is towing a glider B. The glider is
gaining altitude. The tow cable has a length r = 60m, and is increasing at a constant rate of 5 degrees per
second. We want to determine the magnitude of the velocity, v G , and the acceleration, aG , of the glider for
the instant when = 15o .

The velocity of the glider will be v G = v A + v G/A . The velocity of the glider relative to A is best expressed
using a polar coordinate system. Thus, we write,
v G/A = rer + re .
Since the length of the cable is constant, vr = r = 0, and = 5(2/360) = 0.0874 rad/sec, which gives
v = r = 5.236 m/s. We can now transform the velocity vector to cartesian coordinates to obtain,
v G/A = 5.236e = 5.236 sin 15o i + 5.236 cos 15o j = 1.3552i + 5.0575j m/s ,
or,
v G = ((200/3.6) + 1.3552)i + 5.0575j m/s,

vG = 57.135 m/s = 205.7 km/h .

For the acceleration aG = aG/A , since aA = 0. We also know that = 0, and, obviously, r = 0, thus,
2
aG/A = r2 er = 0.457er m/s ,

aG = 0.457 m/s .

Example

Aircraft collision avoidance

Consider two aircrafts, A and B, ying horizontally at the same level with constant velocities v A and v B .
Aircraft B is capable of determining by measurement the relative position and velocity of A relative to B,
i.e., r A/B and v A/B . We want to know whether it is possible to determine from this measurement whether
the two aircraft are on a collision course.

Take the origin O at the hypothetical point of collision where the two trajectories intersect. If collision is to
occur, the time taken for both aircraft to reach O will be the same. Therefore,
rB
rA
=
.
vB
vA
Now consider the velocity triangle formed by v A = v B + v A/B .

Clearly, v A is parallel to r A , and v B is parallel to r B . In addition, if collision is to occur, v A/B will be


parallel to r A/B , since the position and velocity triangles will, in this case, be similar. We see, therefore,
that the condition for the two planes to collide is that the relative velocity between them is parallel to the
relative position vector. If aircraft B uses a polar coordinate system (common for radar measurements),
then the relative position of A with respect to B will be of the form r A/B = rer . Therefore, the velocity
also has to be of that form, which implies that v = r = 0, or = constant. That is, the angle at which B
sees A (sometimes called the bearing angle) should not change.

Example

Velocity made good to windward

In sailboat racing, the most important variable is the component of velocity made by the sailboat in the
direction of the wind. Typically sailboats can sail at an angle of 45o relative to the wind direction. Consider
the sailboat 1 making a velocity vector v1 at an angle of 45o with respect to the wind in the xed coordinate
systemhe knows his angle relative to the wind from compass measurements. He is racing sailboat 2 who
makes a velocity vector v2 at an angle to the wind of 2 , neither of which is known to the skipper of sailboat
1.

The skipper of sailboat 1 would like to know whether he is traveling faster in the direction of the wind
than sailboat 2. He sees a distant lighthouse at an angle of 45o from his direction of travel. He lines up
his opponent with this stationary lighthouse. How can he use this observation to determine whether he is
winning the race to the weather mark?
The instantaneous observation of relative velocity in the direction of the wind can be determined by whether
sailboat 2 pulls ahead of the line between the lighthouse and a xed point on sailboat 1. Sailboat 1 can
observe the sign of V2 cos 2 V1 cos 1 but not these quantities individually. However, that combination,
which he can observe by sighting, is the dierence in the velocities of the two sailboats made good in the
direction of the wind. Therefore, if sailboat 2 drops behind the line of sight to the distant lighthouse, sailboat
1 is winning the race. However, if sailboat 2 pulls ahead of the line of sight to the distant lighthouse, sailboat
2 is winning the race.

ADDITIONAL READING
7

J.L. Meriam and L.G. Kraige, Engineering Mechanics, DYNAMICS, 5th Edition
2/8

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

S. Widnall, J. Peraire
16.07 Dynamics
Fall 2009
Version 2.0

Lecture L8 - Relative Motion using Rotating Axes


In the previous lecture, we related the motion experienced by two observers in relative translational motion
with respect to each other. In this lecture, we will extend this relation to another type of observer. That
is, an observer who rotates relative to a stationary observer. Later in the term, we will consider the more
general case of a translating (includes accelerating) and rotating observer as well as multiple observers who
may be translating and rotating with respect to one another.
Although governed by the same equations, it is useful to distinguish two cases: rst is the rotating observer
that observes a phenomena that a stationary observer would categorize as non-accelerating motion such as a
xed object or one moving at constant velocity (and direction). Such an observer might be a child riding on a
merry-go-round or play-ground turntable. Such a rotating observer would observe a xed object or one with
constant velocity motion as tracing a curved trajectory with changing velocity and therefore accelerating
motion. Our transformation from the rotating into the xed system, must show that the observed motion
was non-accelerating and thus force-free by Newtons law. We should also be able to describe the motion
seen by the rotating observer.
The second case is where a physical phenomena takes place on a platform that is rotating, such as a massspring system xed to a turntable. In this case, the motion, the forces and the frequency of oscillation will
be aected by the rotation and will dier from the system behavior in an inertial coordinate system. Our
approach must demonstrate this dierence and provide the tools for analyzing the system dynamics in the
rotating frame.
As a matter of illustration, let us consider a very simple situation, in which a point a at rest with respect to
the xed observer A located at the origin of the coordinates x, y, z, point O, is also observed by a rotating
observer, B, who is also located at point O. The coordinate system used by B, x , y , z is instantaneously
aligned with x, y, z but rotating with angular velocity . In a), the coordinate system x, y, z and x , y , z
are shown slightly oset for emphasis.

Consider rst case shown in a), in which the point a located at position r does not move in inertial space.
Observer A will observe v a = 0 for the point r. Now consider the same situation observed by B. In the
rotating coordinate system, B will observe that the point r, which is xed in inertial space appears to move
backwards due to the rotation of B s coordinate system.
v a/B = r.

(1)

To reconcile these two observations, it is clear that to obtain the velocity in inertial coordinates, A must
correct for the articial velocity seen by B due to the rotation of B s coordinate system by adding r
to the velocity reported by B.
v a/A = v a/B + r

(2)

In this case, we obtain the rather trivial result that


v a = 0 = r + r.

(3)

If as is shown in b)the point a is moving with velocity v a/A = r A then we have the more general result
v a/A = v a/B + r

(4)

Another way to obtain this expression is directly from Coriolis theorem relating the derivative of a vector
seen by an observer in a rotating coordinate system to that seen in inertial space.
r A = r B + r
In order to apply this relationship, it is not necessary that the vector be constant in time. may be
changing in magnitude or direction.
We have now completed the analysis of the position vector and the velocity vector in a co-located inertial
and rotating coordinate system. To determine the acceleration, we again discuss the general formula for the
time derivative of a vector in a rotating frame, Coriolis Theorem.
2

Time derivative of a xed vector in a rotating frame


We consider the reference frame of x y z rotating with an angular velocity with respect to a xed frame
xyzwith which it is instantaneously aligned. We consider the more general case where is not aligned
with the z axis. Let V be any vector, which is constant relative to the frame x y z . That is, the vector
components in the x y z frame do not change, and, as a consequence, V rotates as if it were rigidly attached
to the frame.

In the absolute frame, the time derivative will be equal to


dV
(V + V ) V
= lim
=V ,
t0
dt
t
which can be interpreted as the velocity of the tip of vector V . The above expression applies to any vector
which is rigidly attached to the frame x y z .

Time derivative of a vector in a rotating frame: Coriolis theorem


Now let V be an arbitrary vector (e.g. position, velocity), which is allowed to change in both the xed xyz
frame and the rotating x y z frame. If we now consider the time derivative of V , as seen by the xed frame,
we have
dV
V xyz =
= (V )x y z + V .
dt

(5)

Here, (V )x y z is the time derivative of the vector V as seen by the rotating frame. The expression (5) above
is known as Coriolis theorem. Given an arbitrary vector, it relates the derivative of that vector as seen by a
xed frame with the derivative of the same vector as seen by a rotating frame. Symbolically, we can write,
( )xyz = ( )x y z + (

).

We will often omit the notation ( )xyz when the derivative is taken with respect to the xed frame.

Summary: Relationships between position, velocity and accelera


tion vectors from rotating to non-rotating coordinate systems
We will now systematically apply Coriolis theorem to determine the relationship between the position,
velocity and acceleration vectors as seen by observer A and B.

Position vector
Since we have dened our coordinate system as having a common origin O and being instantaneously
aligned , the position vector r seen by both the xed observer A and the rotating observer B at this instant
are the same.
r a/A = r a/B

(6)

Note that this is an instantaneous concept, and therefore it is immaterial whether the observer B is rotating
or not. In other words, the above expression is valid at any given instant. Of course, if we choose to express
r a/A in xyz components, and r a/B in x y z components, then we will have to make sure that the proper
coordinate transformation is done before the components of the vectors are added. Also, since the orientation
of x y z changes in time, this transformation will depend on the instant considered.

Velocity vector
Dierentiating (6) with respect to time, we have,
v a/A = r a

(r a/B )x y z + r a/B

(v a/B )x y z + r a/B .

(7)

Here, we have used Coriolis theorem (5) to calculate the derivative of r a/B , i.e. r a/B = (r a/B )x y z +
r a/B . In the above expression, v a/A is the velocity of a relative to the xed frame. The term (v a/B )x y z
is the velocity of a measured by the rotating observer, B. Finally, is the angular velocity of the rotating
frame, and ra/B is the relative position vector of a with respect to B.

Acceleration vector
Dierentiating (7) once again, and making use of Coriolis theorem, e.g.
d(v a/A )/dt = (v a/B ) + (v a/B )

(8)

d(v a/A )/dt = d((dr a/B /dt + r a/B )/dt + (d(r a/B )/dt + r a/B )

(9)

or

we obtain the following expression for the acceleration,


aa/A = v a/A

(v a/B ) + (v a/B )

r a/B + (v a/B ) + ( r a/B )


+
=

r a/B + ( r a/B ) .
(aa/B ) + 2 (v a/B ) +

(10)

The term (aa/B ) is the acceleration of A measured by an observer B that rotates with the axes x y z
instantaneously aligned with x, y, z. The term 2 (v a/B ) is called Coriolis acceleration. The term
r a/B is due to the change in , and, nally, the term ( r a/B ) is called centripetal acceleration.

It can be easily seen that this acceleration always points towards the axis of rotation, and is orthogonal to
(when trying to show this, you may nd the vector identity, A (B C) = (A C)B (A B)C, introduced
in Lecture 2, applied to ( r a/B ) = ( r a/B ) ( )r a/B , useful).

A vs. (
)B

Note

, we did not specify whether the derivative


When writing the derivative of the angular acceleration vector,
was taken with respect to the xed observer, A, or with respect to the rotating observer, B. This may seem
a little bit sloppy at rst, but, in fact, it turns out that it does not really matter. The time derivatives of
vectors which are parallel to are the same for both observers. This can be easily seen if we go back to
Coriolis theorem (5) and apply it to . That is,
A = (
)B + = (
).

since = 0, (cross product).

Acceleration in Rotating Cartesian Coordinates


For later application to problems involving orbital motion (and other applications), we now derive the
expression for acceleration in inertial space seen by the rotating observer B in the cartesian coordinates
shown in a). It is this expression for acceleration which must be used in the application of Newtons Law
F = ma in a rotating coordinate system.
5

Without loss of generality, we take to be aligned with the z, z axis. We have also allowed to be a
function of time although in most applications will be constant. The position vector is r = x i+y j+z k
and = k. The components of acceleration are
ax

ay

az

d2 x
dy
2
d

2
dt2
dt
dt
d2 y
dx
2
d
y +x
+ 2
dt2
dt
dt
d2 z
dt2

(11)
(12)
(13)

Note that since the vector is aligned with the z axis, the rotation of the coordinate system does not
add additional terms to the acceleration in the z direction. The various terms in the x, y equations are
the Coriolis and centripetal accelerations expressed in cartesian coordinates plus the eects of time-varying
. These are the expressions for acceleration that must be used in applying Newtons Law to motion in a
rotating cartesian coordinate system. Of signicant importance is the setting of boundary conditions. We
need 4 boundary conditions: the x y position of the particle; and the x, y components of velocity as seen
by the rotating observer. Since the xed and rotating coordinates are instantaneously aligned in our analysis,
the position seen by both inertial and rotating observers is the same.
To determine the velocity boundary conditions, we distinguish between two cases: a) the observer and b)
the actor.
a) the observer: If the phenomena takes place in inertial space,and the rotating observer simply observes
6

the motion, then the velocity of the particle in inertial space is v a/A = v0 , and the initial position
is r = R0 . In this case, the boundary condition on velocity seen by the rotating observer is must be
corrected to reect the eect of his motion so that in the rotating coordinate system v a/B = v0 R0 .
b) the actor: If the rotating observer actually caused the motion he is observing, such as an astronaut
tossing a cheeseburger to another astronaut, then the boundary condition to be applied to particle
velocity is that observed (or caused) by the actor, v a/B = v0 .

Note

Example: Constant velocity motion observed by an observer rotating at constant

We now consider a particle moving at constant velocity as shown in b) and consider what trajectory the
rotating observer B would observe. We take to be constant in time as well. For a particle moving with
constant velocity, the motion observed by the rotating observer satises equation (9) and (10) with ax = 0
and ay = 0; for simplicity we will consider two-dimensional planar motion.

d2 x
dy
2

2
dt2
dt

d2 y
dx
y 2 + 2
2
dt
dt

(14)
(15)

Of signicant importance is the setting of boundary conditions. We need 4 boundary conditions: the x y
position of the particle; and the x, y components of velocity as seen by the rotating observer. Since the xed
and rotating coordinates are instantaneously aligned in our analysis, the position seen by both inertial and
rotating observers is the same.
These equations have a closed form solution. We consider the simple case shown in b): a particle located at
x = R0 , y = 0 with velocity vx = 0 and vy = v0 . If v0 is less than R0 , the particle observed by B would
actually appear to move backwards along a trajectory as shown in b).

ADDITIONAL READING
J.L. Meriam and L.G. Kraige, Engineering Mechanics, DYNAMICS, 5th Edition
5/7, 7/6

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

J. Peraire, S. Widnall
16.07 Dynamics
Fall 2009
Version 2.0

Lecture L9 - Linear Impulse and Momentum. Collisions


In this lecture, we will consider the equations that result from integrating Newtons second law, F = ma,
in time. This will lead to the principle of linear impulse and momentum. This principle is very useful when
solving problems in which we are interested in determining the global eect of a force acting on a particle
over a time interval.

Linear Momentum
We consider the curvilinear motion of a particle of mass, m, under the inuence of a force F . Assuming that
the mass does not change, we have from Newtons second law,
F = ma = m

dv
d
= (mv) .
dt
dt

The case where the mass of the particle changes with time (e.g. a rocket) will be considered later on in this
course. The linear momentum vector, L, is dened as
L = mv .
Thus, an alternative form of Newtons second law is
,
F =L

(1)

which states that the total force acting on a particle is equal to the time rate of change of its linear momentum.

Principle of Linear Impulse and Momentum


Imagine now that the force considered acts on the particle between time t1 and time t2 . Equation (1) can
then be integrated in time to obtain
t2

t2

F dt =
t1

dt = L2 L1 = L .
L

(2)

t1

Here, L1 = L(t1 ) and L2 = L(t2 ). The term


t2
I=
F dt = L = (mv)2 (mv)1 ,
t1

is called the linear impulse. Thus, the linear impulse on a particle is equal to the linear momentum change
L. In many applications, the focus is on an impulse modeled as a large force acting over a small time. But
1

t
in fact, this restriction is unnecessary. All that is required is to be able to perform the integral t12 F dt. If
t
the force is a constant F, then L = t12 F dt = F (t2 t1 ). If the force is given as a function of time, then
t
L = t12 F (t) dt
Note

Units of Impulse and Momentum

It is obvious that linear impulse and momentum have the same units. In the SI system they are N s or kg
m/s, whereas in the English system they are lb s, or slug ft/s.

Example (MK)

Average Drag Force

The pilot of a 90, 000-lb airplane which is originally ying horizontally at a speed of 400 mph cuts o all
engine power and enters a glide path as shown where = 5o . After 120 s, the airspeed of the plane is 360
mph. We want to calculate the magnitude of the time-averaged drag force.

Aligning the x-axis with the ight path, we can write the x component of equation (2) as follows

120

(W sin D) dt = Lx (120) Lx (0) .


0

, is
The time-averaged value of the drag force, D
= 1
D
120

120

D dt .
0

Therefore,
)120 = m(vx (120) vx (0)) .
(W sin D
Substituting and applying the appropriate unit conversion factors we obtain,
)120 = 90, 000 (360 400) 5280 D
= 9, 210 lb .
(90, 000 sin 5o D
32.2
3600

Impulsive Forces
We typically think of impulsive forces as being forces of very large magnitude that act over a very small
interval of time, but cause a signicant change in the momentum. Examples of impulsive forces are those
generated when a ball is hit by a tennis racquet or a baseball bat, or when a steel ball bounces on a steel
plate. The table below shows typical time intervals over which some of these impulses occur.
Time interval t [s]
Racquet hitting a tennis ball
Bat hitting a baseball

0.005 0.05
0.010.02

Golf club hitting a golf ball


Pile Driver

0.001
0.01 0.02

Shotgun

0.001

Steel ball bouncing on steel plate

0.0002

Example (MK)

Baseball Bat

A baseball is traveling with a horizontal velocity of 85 mph just before impact with the bat. Just after the
impact, the velocity of the 5 18 oz. ball is 130 mph, at an angle of 35o above the horizontal. We want to
determine the horizontal and vertical components of the average force exerted by the bat on the baseball
during the 0.02 s impact.

If we consider the time interval between the instant the baseball hits the bat, t1 , and the instant after it
leaves the bat, t2 , the forces responsible for changing the baseballs momentum are gravity and the contact
forces exerted by the bat. We can use the x and y components of equation (2) to determine the average
force of contact. First, consider the x component,
t2
mvx1 +
Fx dt = mvx2 .
t1

Inserting numbers and using appropriate conversion factors, we obtain

5.125/16
5280
5.125/16
5280
(85
) + Fx (0.02) =
(130
cos 35o ) Fx = 136.7 lb .
3600
32.2
3600
32.2

For the y direction, we have

t2

mvy1 +

Fy dt = mvy2 ,
t1

which gives,
0 + Fy (0.02)

5.125
5.125/16
5280
(0.02) =
(130
sin 35o ) Fy = 54.7 lb .
16
32.2
3600

We note that mgt = 0.0064 lb-s, whereas Fy (t) = 1.094 lb - s. Thus, mgt is 0.59% of the total impulse.
We could have safely neglected it. In this case we would have obtained Fy = 54.4 lb.

Conservation of Linear Momentum


We see from equation (1) that if the resultant force on a particle is zero during an interval of time, then its
linear momentum L must remain constant. Since equation (1) is a vector quantity, we can have situations
in which only some components of the resultant force are zero. For instance, in Cartesian coordinates, if the
resultant force has a non-zero component in the y direction only, then the x and z components of the linear
momentum will be conserved since the force components in x and z are zero.
Consider now two particles, m1 and m2 , which interact during an interval of time. Assume that interaction
forces between them are the only unbalanced forces on the particles. Let F be the interaction force that
particle m2 exerts on particle m1 . Then, according to Newtons third law, the interaction force that particle
m1 exerts on particle m2 will be F . Using expression (2), we will have that L1 = L2 , or L =
L1 + L2 = 0. That is, the changes of momentum of particles m1 and m2 are equal in magnitude and
opposite in sign, and the total momentum change equals zero. Recall that this is true if the only unbalanced
forces on the particles are the interaction forces. The more general situation in which external forces can be
present will be considered in future lectures.
We note that the above argument is also valid in a componentwise sense. That is, when two particles interact
and there are no external unbalanced forces along a given direction, then the total momentum change along
that direction must be zero.
Example

Ballistic Pendulum

The ballistic pendulum is used to measure the velocity of a projectile by observing the maximum angle max
to which the box of sand with the embedded projectile swings. Find an expression that relates the initial
velocity of a projectile v0 of mass m to the maximum angle max reached by the pendulum. The mass of the
sand box is M and the length of the pendulum is L.

We consider the equation of conservation of linear momentum along the horizontal direction. The initial
momentum of the projectile is mv0 . Since the sand box is initially at rest its momentum is zero. Just
after the projectile penetrates into the box, the velocity of the sand box and the projectile are the same.
Therefore, if v1 is the velocity of the sand box (with the embedded projectile) just after impact, we have
from conservation of momentum,
mv0 = (M + m)v1 .
After impact, the problem reduces to that of a simple pendulum. The only force doing any work is gravity and
therefore we can apply the principle of conservation of work and energy. At the point when is maximum,
the velocity will be zero. From energy conservation we nally obtain,
1
(M + m)v12 = (M + m)ghmax ,
2
and since h = L(1 cos ), we have

max = cos

m
M +m

v02
2Lg

Note that energy is not conserved in the collision. The initial kinetic energy of the system is the kinetic
energy of the projectile T0 =

1
2
2 mv0

(taking the reference height as zero). After the collision the kinetic

energy of the system is


T1 =

1
(m + M )v12 .
2

(3)

T1 =

1
m2
v2
2 (M + m) 0

(4)

m
Since v1 = v0 (m+M
) we have

which is less than T0 for M > 0.

Collisions
We consider now, the situation of two isolated particles colliding. The only forces on the particles are due
to their mutual interaction. When the velocity vector of the two particles is parallel to the line joining the
two particles, we say that we have one-dimensional collisions, otherwise we say that the collision is oblique.
5

1D Collisions
Here, we are dealing with rectilinear motion and therefore the velocity vector becomes a scalar quantity. In
order to understand the collision process we consider ve dierent stages.
I) Before Impact
Let particle 1, of mass m1 , occupy position x1 and travel with velocity v1 along the direction parallel
to the line joining the two particles. And let particle 2, of mass m2 , occupy position x2 and travel with
speed v2 also in the same direction. We assume that v1 > v2 so that collision will occur.

We can introduce the position of the center of mass xG = (m1 x1 + m2 x2 )/m, where m = m1 + m2 . The
velocity of the center of mass is then given by vG = (m1 v1 + m2 v2 )/m. Note that since conservation
of momentum requires (m1 v1 + m2 v2 ) = (m1 v1 + m2 v2 ), the velocity of the center of mass remains
unchanged by the collision process between the particles. If we dene the relative velocity g = v1 v2
we can express v1 and v2 as a function of vG and g as,
v1
v2

m2
g
m
m1
g.
= vG
m
= vG +

II) Deformation
The particles establish contact and the force between them Fd increases until the instant of maximum
deformation.

III) Maximum Deformation


The contact force is at its maximum and the two particles travel at the same velocity vG . Thus the
deformation force has slowed m1 down to a velocity of vG and sped up m2 to a velocity of vG .
The impulse applied to particle 1 and 2 from the deformation force equals the change in momentum
in the deformation process.
6

For particle 1

Fd dt = m1 vG m1 v1

(deformation phase)

and for particle 2

Fd dt

= m2 vG m2 v2

(deformation phase)

IV) Restoration

The contact force Fr decreases and the particles move apart.

The impulse applied to particle 1 and 2 from the restoring force equals the change in momentum in
the restoration process. After the restoration process the velocity of m1 is v1 and the velocity of m2 is
v2 . For particle 1

Fr dt

m1 v1 m1 vG

(restoration phase)

and for particle 2

Fr dt =

m2 v2 m2 vG

V) After Impact
The particles travel with a constant velocity v1 and v2 .

(restoration phase)

From momentum conservation, the total momentum before and after impact should remain the same,
i.e. m1 v1 +m2 v2 = m1 v1 +m2 v2 , and therefore the velocity of the center of mass will remain unchanged

vG
= vG , thus mvG = m1 v1 + m2 v2 . If g = v1 v2 is the relative velocity of the two particles after

impact, then
v1
v2

m2
g
m
m1
= vG
g .
m
= vG +

(5)
(6)

Coecient of Restitution
We dene the coecient of restitution as the ratio between the restoration and deformation impulses.

Fr dt
e=
Fd dt
For physically acceptable collisions 0 < e < 1. The value of e = 1 corresponds to an elastic collision, whereas
the value of e = 0 corresponds to a totally inelastic collision in which the restoration impulse is equal to
zero.
We can consider each particle separately and set the impulse on the particle equal to the change of linear
momentum
Since for particle 1

Fd dt

m1 vG m1 v1

(deformation phase)

Fr dt

m1 v1 m1 vG

(restoration phase)

Thus, e = (v1 vG )/(vG v1 ), or (e + 1)vG = v1 + ev1 .


and for particle 2

Fd dt =

m2 vG m2 v2

(deformation phase)

Fr dt =

m2 v2 m2 vG

(restoration phase)

Thus, e = (v2 vG )/(vG v2 ), or (e + 1)vG = v2 + ev2 .


Combining the above expressions, we have

Fr dt
=

e=
Fd dt

v1 v2
g
= ,
v1 v2
g

which expresses the coecient of restitution as the ratio of the relative velocities after and before the collision.
Finally, if we know e we can use equations (5) and (6) to obtain

v1
v2

m2
eg
m
m1
= vG +
eg .
m
= vG

Example

Simple Collisions

As an example, we consider a simple case of a mass m2 at rest and a mass of m1 traveling at velocity v1
headed for a collision with m2 . After the collision, the velocities of the two masses will in general be non-aero.
We consider the cases e = 1, .5, and0, the is a range of coecient of restitution between 1, perfectly elastic,
to zero, perfectly inelastic. The result for the nal velocities of m1 and m2 , normalized by the initial velocity
v1 , are given by the equations above and shown in the gure.

For e = 0, these is only one curve since the particles stick together: v1 /v1 = v2 /v1 . For the elastic collision,
note the result for equal masses: v1 = 0 and v2 /v1 = 1

Energy
We can now look at the kinetic energy before and after the impact. The kinetic energy before impact is,
T

=
=
=

1
1
m1 v12 + m2 v22
2
2
1
m2
m2
1
m1
m2
2
2
m1 (vG
+2
gvG + 22 g 2 ) + m2 (vG
2
gvG + 12 g 2 )
2
m
m
2
m
m
1
1 m1 m2
1
1 2
2
2
2
mv +
(v1 v2 ) = mvG + g .
2 G 2 m
2
2

Here we introduce the notation = (m1 m2 )/m; this is often referred to as the reduced mass of the pair. The
reduced mass approaches 1/2 of each mass when the two masses are similar, but it approaches the lighter
mass if the masses are dissimilar. The kinetic energy after impact will be
T

=
=
=

since g = eg and

m1 m2 )

1
mv 2 +
2 G
1
mv 2 +
2 G
1
mv 2 +
2 G

1 m1 m2
(v1 v2 )2
2 m
1 m1 m2 2
g
2 m
1 2 2
e g .
2

= . Therefore, the kinetic energy lost during the collision is


E = T T =

1
(1 e2 )g 2 ,
2

which is zero for e = 1 (elastic collision), and maximum when e = 0.


Momentum exchange
We can also look at the momentum which is exchanged between particle 1 and particle 2,
M12 = m1 v1 m1 v1 = m1 (vG +

m2
m2
g vG +
eg) = (1 + e)g.
m
m

This, is maximum when e = 1 (elastic collision) and minimum when e = 0.


It is particularly illuminating to examine a special case of collisions, that for which the two masses are equal.

We consider two cases shown in the gure. In both cases one particle is at rest, the other approaching with
velocity v1 . In one case, we assume an elastic collision, e = 1; in the other a completely inelastic collision,
10

e = 0. In the rst case, kinetic energy is conserved so it is obviousalthough you can work out the algebra
that the rst particle completely transfers its momentum to the second particle, where upon it continues
to the right with the original velocity v2 = v1 . In the second case of a completely inelastic collision, there
is no rebounding force and the particles stick and move together. Conservation of momentum requires the
v2 = v1 /2. Kinetic energy is not conserved.
Example

Bouncing Ball

If a ball, initially at rest, is released on a at surface from a height h and it rebounds to a height h , the
coecient of restitution is given by
v
e= =
v

h
,
h

where v and v are the velocities before and after impact.

Oblique Collisions
In the case of oblique collisions, we consider the instant of impact and dene the normal direction, n, along
the line that connects the two mass centers, and the tangential direction, t, along the line tangent to the
surfaces at the point of contact.

We write the equations of conservation of momentum in the tangent and normal directions. Since the force
of contact is assumed to act along the normal direction, the conservation of linear momentum along the
tangential direction implies
m1 (v1 )t = m1 (v1 )t

(v1 )t = (v1 )t

m2 (v2 )t = m2 (v2 )t

(v2 )t = (v2 )t

In the normal direction, we solve a 1D collision problem, that is,


= m1 (v1 )n + m2 (v2 )n
(v )n (v2 )n
e = 1
(v1 )n (v2 )n

m1 (v1 )n + m2 (v2 )n

which determine (v1 )n and (v2 )n .


11

ADDITIONAL READING
J.L. Meriam and L.G. Kraige, Engineering Mechanics, DYNAMICS, 5th Edition
3/8, 3/9

12

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

S. Widnall, J. Peraire
16.07 Dynamics
Fall 2008
Version 2.0

Lecture L10 - Angular Impulse and Momentum for a Particle


In addition to the equations of linear impulse and momentum considered in the previous lecture, there is a
parallel set of equations that relate the angular impulse and momentum.

Angular Momentum
We consider a particle of mass m, with velocity v, moving under the inuence of a force F . The angular
momentum about point O is dened as the moment of the particles linear momentum, L, about O.

Thus, the particles angular momentum is given by,


H O = r mv = r L .

(1)

The units for the angular momentum are kgm2 /s in the SI system, and slugft2 /s in the English system.

It is clear from its denition that the angular momentum is a vector which is perpendicular to the plane
dened by r and v. Thus, on some occasions it may be more convenient to determine the direction of H O
from the right hand rule, and its modulus directly from the denition of the vector product,
HO = mvr sin ,
where is the angle between r and v.

In other situations, it may be convenient to directly calculate the angular momentum in component form.

For instance, using a right handed cartesian coordinate system, the components of the angular momentum

are calculated as

HO

= Hx i + Hy j + Hz k = x

mvx

j
y
mvy

z = m(vz y vy z)i + m(vx z vz x)j + m(vy x vx y)k .

mvz
k

Similarly, in cylindrical coordinates we have

er
e

H O = Hr er + H e + Hz k = r
0

mvr mv

z = mv zer + m(vr z vz r)e + mv rk .

mvz
k

Rate of Change of Angular Momentum


We now want to examine how the angular momentum changes with time. We examine this in two dierent
coordinate systems: system a) is about a xed point O; system b) is about the center of mass of the particle.
Of course system b) is rather trivial for a point mass, but its later extensions to nite bodies will be extremely
important. Even at this trivial level, we will obtain an important result.

About a Fixed Point O


The angular momentum about the xed point O is
HO = r mv
2

(2)

Taking a time derivative of this expression , we have

O = r mv + r mv .

H
Here, we have assumed that m is constant. If O is a xed point, then r = v and r mv = 0. Thus, we end
up with,
O = r mv = r ma .
H
Applying Newtons second law to the right hand side of the above equation, we have that r ma = r F =
M O , where M O is the moment of the force F about point O. The equation expressing the rate of change
of angular momentum is then written as
O .
MO = H

(3)

We note that this expression is valid whenever point O is xed. The above equation is analogous to the
equation derived in the previous lecture expressing the rate of change of linear momentum. It states that
the rate of change of linear momentum about a xed point O is equal to the moment about O due to the
resultant force acting on the particle. Since this is a vector equation, it must be satised for each component
independently. Thus, if the force acting on a particle is such that the component of its moment along a given
direction is zero, then the component of the angular momentum along this direction will remain constant.
This equation is a direct consequence of Newtons law. It will not give us more information about the
momentum of a particle, but a clever choice of coordinates may make angular momentum easier to apply in
any given case.

About the Center of Mass


For a particle, the angular momentum is zero. We examine carefully the expression for the rate of change of
angular momentum.
O = r mv + r mv .
H
Since the coordinate system moves with the particle, both r and v are zero. This is true even if the coordinate
system is not inertial, in contrast to the application of Newtons Law for linear momentum. Therefore, for a
mass point the rate of change of angular momentum is zero in a coordinate system moving with the particle.
This implies that no moments can be applied to a mass point in a coordinate system moving with the point.
This result will later be extended to bodies of nite size; we can apply Equation 3 to a body of nite size
even in an accelerating coordinate system if the origin of our coordinate system is the center of mass.
Example

Pendulum

Here, we consider the simple pendulum problem. However, we assume that the point mass m is suspended
from a rod attached at a pivot that could support a side force. Therefore we allow the direction of the force
to be unknown. We rst apply conservation of angular momentum in a coordinate system moving with the
3

point mass, the x , y system. The result that the angular momentum in this coordinate system is zero, gives
us an immediate result that no external torques can act on the particle. Gravity acts at the particle and
therefore produces no moment. Therefore we conclude that the force in the rod must point directly to the
mass, along the rod itself. In other words the rod acts as a string supporting a tension T. We will later see
that if the pendulum has a nite moment of inertia about the center of mass, this result no longer applies.
We now examine the pendulum in a coordinate system xed at the point O and re-derive the pendulum
equation using equation (3).

There are two forces acting on the suspended mass: the string tension and the weight. By our earlier
argument, the tension, T , is parallel to the position vector r and therefore its moment about O is zero. On
the other hand, the weight creates a moment about O which is M O = lmg sin k.
The angular momentum is given by
H O = r mv = ler mle = ml2 er e = ml2 k.
Therefore, the z component of equation (3) gives
ml2 = lmg sin ,
or,
g
+ sin = 0 ,
l
which is precisely the same equation as the one derived in lecture L5 using Newtons law. The derivation
using angular momentum is more compact.

Principle of Angular Impulse and Momentum


Equation (3) gives us the instantaneous relation between the moment and the time rate of change of angular
momentum. Imagine now that the force considered acts on a particle between time t1 and time t2 . Equation
(3) can then be integrated in time to obtain
t2
t2
O dt = (H O )2 (H O )1 = H O .
M O dt =
H
t1

(4)

t1

Here, (H O )1 = H O (t1 ) and (H O )2 = H O (t2 ). The term


t2
M O dt ,
t1

is called the angular impulse. Thus, the angular impulse on a particle is equal to the angular momentum
change.
Equation (4) is particularly useful when we are dealing with impulsive forces. In such cases, it is often
possible to calculate the integrated eect of a force on a particle without knowing in detail the actual value
of the force as a function of time.

Conservation of Angular Momentum


We see from equation (1) that if the moment of the resultant force on a particle is zero during an interval of
time, then its angular momentum H O must remain constant.
Consider now two particles m1 and m2 which interact during an interval of time. Assume that interaction
forces between them are the only unbalanced forces on the particles that have a non-zero moment about a
xed point O. Let F be the interaction force that particle m2 exerts on particle m1 . Then, according to
Newtons third law, the interaction force that particle m1 exerts on particle m2 will be F . Using expression
(4), we will have that (H O )1 = (H O )2 , or H O = (H O )1 + (H O )2 = 0. That is, the changes
in angular momentum of particles m1 and m2 are equal in magnitude and of opposite sign, and the total
angular momentum change equals zero. Recall that this is true only if the unbalanced forces, those with
non-zero moment about O, are the interaction forces between the particles. The more general situation in
which external forces can be present will be considered in future lectures.
We note that the above argument is also valid in a componentwise sense. That is, when two particles interact
and there are no external unbalanced moments along a given direction, then the total angular momentum
change along that direction must be zero.
Example

Ball on a cylinder

A particle of mass m is released on the smooth inside wall of an open cylindrical surface with a velocity v 0
that makes an angle with the horizontal tangent. The gravity acceleration is pointing downwards. We
want to obtain : i) an expression for the largest magnitude of v 0 that will prevent the particle from leaving
5

the cylinder through the top end, and ii) an expression for the angle that the velocity vector will form
with the horizontal tangent, as a function of b.

The only forces on the particle are gravity and the normal force from the cylinder surface. The moment of
these forces about O (or, in fact, about any point on the axis of the cylinder) always has a zero component in
the z direction. That is, (MO )z = 0. To see that, we notice that for any point on the surface of the cylinder,
r and F are always contained in a vertical plane that contains the z axis. Therefore, the moment must be
normal to that plane. Since the moment has zero component in the z direction, (HO )z will be constant.
Thus, we have that
(HO )z = rmv0 cos = constant .
For part i), we consider the trajectory for which the velocity is horizontal when z = a and let (v 0 )limit be
the initial velocity that corresponds to this trajectory. It is clear that for any trajectory for which v 0 has a
larger magnitude than (v 0 )limit , the particle will leave the cylinder through the top end. Thus, for the limit
trajectory we have, from conservation of energy
1
1
m(v0 )2limit = mva2 + mga ,
2
2
and from conservation of angular momentum
(HO )z = rm(v0 )limit cos = rmva .
Here, va is the magnitude of the velocity for the limit trajectory when z = a. Eliminating va from these
equations we nally arrive at,

(v0 )limit =

2ga
.
1 cos2

Therefore, for v0 (v0 )limit the mass will not leave the cylinder through the top end.

For part ii), we also consider conservation of energy


1
1
mv02 = mvb2 mgb
2
2
and conservation of angular momentum,
rmv0 cos = rmvb cos .
Eliminating, vb from these two expressions we obtain,

= cos

cos

1 + 2gb/v02

Example

Spinning Mass

A small particle of mass m and its restraining cord are spinning with an angular velocity on the horizontal
surface of a smooth disk, shown in section. As the force F s is slowly increased, r decreases and changes.
Initially, the mass is spinning with 0 and r0 . Determine : i) an expression for as a function of r, and ii)
the work done on the particle by F s between r0 and an arbitrary r. Verify the principle of work and energy.

The component of the moment of the forces acting on the particle is zero along the spinning axis. Therefore,
the vertical component of the angular momentum will be constant. For i), we have
mr0 v0 = mrv,

v0 = 0 r0 , v = r

r02 0
.
r2

For ii), we rst calculate the force on the string


Fs = m

v2
r2 2
r4 2
= m
= m 0 3 0 .
r
r
r

The work done by Fs , will be

W =
r0

Fs dr = mr04 02

r
r0

dr
1
= mr04 02
r3
2

1
1
2
r2
r0

The energy balance implies that


T0 + W = T .
This expression can be directly veried since,
2

1
1
r0
1
2
2
m(0 r0 ) + m(0 r0 )
1 = m(r)2 .
2
2
2
r
2


T0

Example

Ballistic Pendulum

We consider a pendulum consisting of a mass, M , suspended by a rigid rod of length L. The pendulum is
initially at rest and the mass of the rod can be neglected. A bullet of mass m and velocity v 0 impacts M
and stays embedded in it. We want to nd out the angle max reached by the pendulum. The angle that
the velocity vector v 0 forms with the horizontal is .

Because the rod is assumed to be rigid, we can expect that when the bullet impacts the mass, there will
be an impulsive reaction that the rod will exert on the bullet. If we use the principle of linear impulse and
momentum, it will be necessary to solve for this impulsive force. An alternative approach that simplies the
problem considerably is to use the principle of angular impulse and momentum. We consider the angular
momentum about point O of the particles m and M just before and after the impact. The only external
forces acting on the two particles are gravity and the reaction from the rod. It turns out that gravity is not
an impulsive force and therefore its eect on the total angular impulse, over a very short time interval, can
be safely neglected (it turns out that in this case, the moment about O of the gravity forces at the time of
impact is also zero). On the other hand, we can expect the reaction from the rod to be large. However, the
moment about O of this reaction is zero, since it is directed in the direction of the rod. Therefore, we have
that during impact, the z component of the angular momentum is conserved.
8

The angular momentum before impact will be,


[(HO )z ]1 = L cos mv0 .
After impact, the velocity v 1 has to be horizontal. Thus, the angular momentum will be
[(HO )z ]2 = L(M + m)v1 .
Equating these two expressions we get,
v1 =

m
v0 cos .
M +m

After impact, the system is conservative, and the maximum height can be easily obtained from conservation
of energy,
1
(M + m)v12 = (M + m)gL(1 cos max ) .
2
Thus,

max = cos

m
(M + m)

v02 cos2
2gL

ADDITIONAL READING
J.L. Meriam and L.G. Kraige, Engineering Mechanics, DYNAMICS, 5th Edition
3/10

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

J. Peraire, S. Widnall
16.07 Dynamics
Fall 2008
Version 2.0

Lecture L11 - Conservation Laws for Systems of Particles


In this lecture, we will revisit the application of Newtons second law to a system of particles and derive
some useful relationships expressing the conservation of angular momentum. We also specialize these results
to two-dimensional rigid bodies.

Center of Mass
Consider a system made up of n particles. A typical particle, i, has mass mi , and, at the instant considered,
occupies the position r i relative to a frame xyz. We can then dene the center of mass, G, as the point
whose position vector, r G , is such that,
rG =

n
1
(
mi r i ) .
m i=1

Here, m is the total mass of the system given by m =

(1)

mi .

i=1

It is important to note that the center of mass is a property of the system and does not depend on the
reference frame used. In particular, if we change the location of the origin O, r G will change, but the
absolute position of the point G within the system will not. Often, it will be convenient to describe the
motion of particle i as the motion of G plus the motion of i relative to G. To this end, we introduce the
relative position vector, r i , and write,
r i = r G + r i .

(2)

It follows immediately, from the denition of the center of mass (1) and the denition of the relative vector
r i (2), that,

i=1

mi r i =

mi (r i r G ) =

i=1

i=1

This result will simplify our later analysis.


1

mi r i mr G = 0 .

(3)

Forces
In order to derive conservation laws for our system, we isolate it a little more carefully, identify what mass
particles it contains and what forces act upon the individual particles.

We will consider two types of forces acting on the particles :


External forces arising outside the system. We will denote the resultant of all the external forces
N
acting on the system as i=1 F Ei = F .
Internal forces due to pairwise particle interactions. Let f ij denote the force that particle j exerts
on particle i directed along the line joining the two particles i and j. This force could arise from
gravitation attraction or from internal force due to the connections between particles. It could also
arise from collisions between individual particles that, as we have seen, produce equal and opposite
impulsive forces that conserve momentum.
By Newtons third law, these internal forces act in pairwise equal and opposite directions. Therefore,
f ij = f ji , where f ji is the force that particle i exerts on particle j along the line joining the particles The
total internal force on particle i is then
n

f ij ,

j=1, j=i

and, if we sum over all particles, we have


n

f ij = 0 ,

i=1 j=1 j=i

since, for every force, f ij , there is an equal and opposite force, f ji for both normal and tangential forces.(The
notation j =
i means to exclude j=i in the sum since of course the particle induces no net force upon itself.)

Conservation of Linear Momentum


The linear momentum of the system is dened as,
L=

mi v i .

(4)

i=1

From equation (2), we have that v i = r i = r G + r i , which, combined with the above equation, gives,
L=

mi (r G + r i ) =

i=1

mi v G +

i=1

since by the denition of center of mass

i=1

n
d
(
mi r i ) = mv G ,
dt i=1

(5)

mi r i = 0. We now consider the time variation of the linear

momentum. If we assume that the reference frame xyz is inertial, then, starting from equation (4), we have,
=
L

i=1

mi ai =

n
n
n

(F i +
f ij ) =
Fi = F,
i=1

j=1 j=i

(6)

i=1

where F is the sum of all external forces acting on the system. Since the sum of the internal forces balance
when summed over i and j, we are left with only the summation over the external forces. Thus, for a system
of particles, we have that,
=F .
L

(7)

= maG . This is a powerful result. Note that the center


Note that, from equation (5), we can also write L
of mass is in general not xed to a particular particle but is a point in space about which the individual
particles move.
These ideas also describe the conservation of linear momentum under external and internal collisions. Since
individual internal collisions between particles in the system conserve momentum, the sum of their interac
tions also conserves momentum. If we consider an external particle imparting momentum to the system, it
could be treated as an external impulse. Conversely we can consider the particle about to collide to be a
part of the system, and include its momentum as part of total system momentum, which is then conserved
by Newtons law.

Conservation of Angular Momentum


Since the angular momentum is dened with respect to a point in space, we will consider two cases, using
a dierent reference point for each case: 1) conservation of angular momentum about a xed (or more
generally a non-accelerating) point O; and 2) conservation of angular momentum about the center of mass,
which could be accelerating in inertial space.
The gure shows the system: the coordinate system x, y, z about the xed point O, the coordinate system
x , y , z about the center of mass, G, the internal forces between particles, the angular momentum H 0 about
the point O and the angular momentum H G about the center of mass, point G, which in general are not
equal.
3

Conservation of Angular Momentum about a Fixed Point O

The angular momentum of a system of particles about a xed point, O, is the sum of the angular momentum
of the individual particles,
HO =

(r i mi v i ) .

(8)

i=1

The time variation of H O can be written as,


O=
H

(ri mi v i ) +

i=1

(r i mi vi ) = 0 +

i=1

(r i (F i +

f ij )) =

j=1, j=i

i=1

n
n

(r i F i ) +
Mi . (9)
i=1

i=1

where we replace mv i by the sum of the forces acting on particle i: mv i = (F i +

j=1, j=i

.f ij ).

i=1

Mi

is the sum of any external moments that act on the system. The term (ri mi v i ) in equation (9) is zero;
since ri = v i , the two vector are parallel and their cross-product is zero. In the second term we may write
r i f ij + r j f ji = (r i r j ) f ij = 0 since the forces are aligned with (r i -r j ), and their values are equal
and opposite, their cross product with (r i -r j ) is zero, and therefore, the internal forces have no net eect
on the total angular momentum change of the particle system. Therefore
O=
H

(r i mi vi ) =

i=1

(r i F i ) +

i=1

Mi .

(10)

i=1

When evaluating the moments which act to change the angular momentum from equation (13), we see that
the sum of the internal moments is zero so that the only moment which acts to change the angular momentum
is the moments created by the external forces about the point O plus any external moments applied to the
system.
Thus, we have that
O = MO ,
H
where M O =

i=1 (r i

F i) +

i=1

(11)

Mi is the total moment, about O, due to the applied external forces

plus any external moments.


4

Conservation of Angular Momentum about G


The angular momentum about the center of mass G is given by,
HG =

(r i mi v i ) .

(12)

i=1

Taking the time derivative of equation (2), we obtain


v i = r i = r G + r i = v G + v i .

(13)

Inserting this expression into equation 12, we obtain


HG =

(r i mi (r G + r i )) =

i=1

since

i=1 (r i

n
n
n

(r i mi r G ) +
(r i mi r i ) =
(r i mi v i ) ,
i=1

mi r G ) = r G

i=1

i=1

(14)

i=1

mi r i = 0 (see equation 3). We note that equations (12) and (14)

give us alternative representations for H G . Equation (12) is called the absolute angular momentum (since
it involves absolute velocities, v i ), whereas equation (14) is called the relative angular momentum (since it
involves velocities, v i , relative to G). When G is chosen to be the origin for the relative velocities, both
the absolute and relative angular momentum are identical. In general, the absolute and relative angular
momentum with respect to an arbitrary point are not the same.
We can now go back to equation (12) and consider the time variation of H G ,
G
H

n
n
n
n

(r i mi (v G + r i )) +
(r i mi v i ) = 0 +
(r i F i ) +
Mi .
i=1

i=1

i=1

In the above equation, the term r i mi r i is clearly zero, and


n
v G d( i=1 mi r i )/dt = 0. Thus, we have that
G=
H

i
i=1 (r

mi v G ) = v G

(r i mi v i ) = M G .

(15)

i=1

i=1

mi r i =

(16)

i=1

Here, M G =

i=1 (r i

F i) +

i=1

Mi , is the total moment, about G, of the applied external forces plus

any external moments. Note that external forces in general produce unequal moments about O and G while
applied external moments (torques) produce the same moment about O and G.
The above expression is very powerful and allows us to solve, with great simplicity, a large class of problems
in rigid body dynamics. Its power lies in the fact that it is applicable in very general situations: In the
derivation of equation (16), we have made no assumptions about the motion of the center of mass, G. That
is, equation (16) is valid even when G is accelerated. We have implicitly assumed that the reference frame
used to describe r i in equation 13 is non-rotating with respect to the xed frame xyz (otherwise, we would
have written r i = v i + r i , with , the angular velocity of the frame considered). It is not dicult to
show that equation (16) is still valid if the reference frame rotates, provided the angular velocity is constant.
If the reference frame rotates with a constant angular velocity, the angular momentum will dier from that
of equations (12) and (14) by a constant, but equation 16 still will be valid.
5

Finally, by combining equations 30 and 12, the angular momentum about a xed point, O, can be expressed
as a function of the angular momentum about the center of mass, as,
H O = r G mv G + H G .

(17)

Just as we could incorporate collisions in our statement of conservation of linear momentum, we can in
corporate collision in our statement of conservation of angular momentum. Collisions conserve both linear
and angular momentum. Just as changes in linear momentum result for linear impulses, changes in angular
momentum result from angular impulses.

Kinetic Energy for Systems of Particles


Here, we derive the expression for the kinetic energy of a system of particles that will be used in the
following lectures. A typical particle, i, will have a mass mi , an absolute velocity v i , and a kinetic energy
Ti = (1/2)mi v i v i = (1/2)mi vi2 .
The total kinetic energy of the system, T , is simply the sum of the kinetic energies for each particle,
T =

Ti =

i=1

1
i=1

mi vi2 .

It is convenient to decompose the velocity of each particle, v i , into the velocity of the center of mass, v G ,
and the velocity relative to the center of mass, r i . Then,
T =

1
i=1

since v G

i=1

mi (v G + r i ) (v G + r i ) =

mi r i = 0, and

1
i=1

i=1

2
mi (vG
+ 2v G r i + ri 2 ) =

1
1
2
mvG
+
mi ri 2 ,
2
2
i=1

mi = m. Thus, we see that the kinetic energy of a system of particles

equals the kinetic energy of a particle of mass total m moving with the velocity of the center of mass, plus
the kinetic energy due to the motion of the particles relative to the center of mass, G. We have said nothing
about the conservation of energy for a system of particles. As we shall see, that depends upon the details of
internal interactions and the work done by the external forces.
We will now particularize the conservation principles presented in the previous lectures to the case in which
the system of particles considered is a 2D rigid body.

Conservation of Angular Momentum for 2D Rigid Body


In subsequent Lectures, we will apply conservation of angular momentum for a general system in three
dimensions. However, the dynamics of a 2D rigid body can easily be incorporated into our study of particle
motion. We therefore specialize these results to the case of a 2D rigid body. The equations describing the
general motion of a rigid body follow from the conservation laws for systems of particles. Since the general
motion of a 2D rigid body can be determined by three parameters (e.g. x and y coordinates of position,
6

and a rotation angle ), we will need to supply three equations. Conservation of linear momentum yields one
vector equation, or two scalar equations. The additional condition is conservation of angular momentum.
We saw that there are several ways to express conservation of angular momentum. In principle, they are all
equivalent, but, depending on the problem situation, the use of a particular form may greatly simplify the
problem. The best choices for the origin of coordinates are: 1) the center of mass G; 2) a xed point O.

Conservation of Angular Momentum about the Center of Mass


When considering a 2D rigid body, the velocity of any point relative to G consists of a pure rotation and,
therefore, can be expressed as
v i = r i ,
where is the angular velocity vector perpendicular to the plane of motion. Both v i and r are in the plane
of the motion. These two equations can be combined to give,
HG =

(r i mi ( r i )) =

i=1

i=1

mi ri 2 =

r2 dm

(18)

For a continuous body, the sum over the mass points is replaced by an integral.

r2 dm is dened as the

mass moment of inertia IG about the center of mass. Here, we have used the vector identity, A (B C) =
(A C)B (A B)C, and imposed the fact that r i and are perpendicular for 2D planar bodies. Thus,
for a 2-D rigid body, the conservation law for angular momentum about the center of mass, G is
IG = M G ,
where IG =

(19)

r2 dm, = and MG is the total moment about G due to external forces and external

moments. Although equation (7) is a vector equation, and M G are always perpendicular to the plane
of motion, and, therefore, equation (7) only yields one scalar equation. The moment of inertia, IG , can
be interpreted as a measure of the bodys resistance to changing its angular velocity as a result of applied
external moments. The moment of inertia, IG , is a scalar quantity. It is a property of the solid which
indicates the way in which the mass of the solid is distributed relative to the center of mass. For example, if
most of the mass is far away from the center of mass, ri will be large, resulting in a large moment of inertia.
The dimensions of the moment of inertia are [M ][L2 ].

Conservation of Angular Momentum about a xed point O


If the xed point O is chosen as the origin, a similar result is obtained. Since for a 2D rigid body the velocity
in the coordinate system xed at the point O is
vi = ri ,

conservation of angular momentum gives


HO =

(r i mi ( r i )) =

i=1

mi ri 2 = ri 2 =

r2 dm = IO

(20)

i=1

where IO is the moment of inertia about the point O. In general, IO =


IG . The conservation law for angular
momentum about the xed point O is then
IO = M O ,
where IO =

(21)

r2 dm, = and MO is the total applied moment due to external forces and moments

(torques). Also it is important to point out that both the angular velocity and the angular acceleration

are the same for any point on a rigid body: G = O , G = O .

Most textbooks on dynamics have tables of moments of inertia for various common shapes: cylinders, bars,

plates. See Meriam and Kraige, Engineering Mechanics, DYNAMICS (Appendix B) for more examples.

Radius of Gyration
It is common to report the moment of inertia of a rigid body in terms of the radius of gyration, k. This is
dened as

I
,
m
and can be interpreted as the root-mean-square of the mass element distances from the axis of rotation.
k=

Since the moment of inertia depends upon the choice of axis, the radius of gyration also depends upon the
choice of axis. Thus we write

IG
,
m
for the radius of gyration about the center of mass, and

IO
kO =
,
m
kG =

for the radius of gyration about the xed point O.

Parallel Axis Theorem


We will often need to nd the moment of inertia with respect to a point other than the center of mass. For
instance, the moment of inertia with respect to a given point, O, is dened as

IO =
r2 dm .

Assuming that O is a xed point, H O = IO . If we know IG , then the moment of inertia with respect

to point O, can be computed easily using the parallel axis theorem. Given the relations r2 = r r and

r = r G + r ., we can then write,

2
2
IO =
r2 dm =
(rG
+ 2r G r + r2 ) dm = rG
m

dm + 2r G
m

r dm +

2
r2 dm = mrG
+ IG ,

since

r dm = 0.

From this expression, it also follows that the moment of inertia with respect to an arbitrary point is mini
mum when the point coincides with G. Hence, the minimum value for the moment of inertia is IG .

Equations of Motion for a 2D rigid Body.


Now that we have developed the equation governing conservation of angular momentum for a 2D rigid body
in planar motion, we can state the governing equations for this three degree of freedom system.
The conservation of linear momentum yields the vector equation,
maG = F ,

(22)

where m is the body mass, aG is the acceleration of the center of mass, and F is the sum of the external
forces acting on the body.
Conservation of angular momentum requires
G = M G = IG = IG
H

(23)

where IG is the moment of inertia about the center of mass; is the angular velocity, whose vector direction
is perpendicular to the x, y coordinate system; and a is the angular acceleration. For impulsive forces, we
write that the change in angular momentum is equal to the time integral of the applied moments, whether
or not these moments are impulsive (but we have to be able to do the time integral, so impulsive forces are
easier.
A body xed at a point O is a single degree of freedom system. Therefore, only one equation is required,
conservation of angular momentum about the point O.
O = M O = IO = IO
H

(24)

Impulsive Motion
Just as we applied the continuous from of Newtons law for a particle to the motion resulting from impulsive
forces and collisions, we can extend these result to impulsive linear and angular motion resulting from
impulsive forces and moments.

t2

H G2 H G1 =
t1

where

(r i F i ) .

(25)

i=1

t2 n
( i=1 r i F i ) = AI is the angular impulse, the integral over time of the sum of the moments
t1

acting about the center of mass. A similar result holds for moments taken about the point O.
t2
n
H O2 H O1 =
(r i F i ) .
t1

i=1

(26)

where

t2 n
( i=1 r i F i ) = AI
t1

Kinetic Energy for a 2D Rigid Body


We specialize the result of the analysis of kinetic energy for a system of particles about the center of mass
for the case of a two-dimensional rigid body for which the velocity of a particle is given by
v i = r i .

(27)

Applying this to the general form for from Equation

T =

1
i=1

mi (v G + r i ) (v G + r i ) =

1
1
mv 2 +
2 G 2

Example

r2 dm =

1
1
mv 2 + 2 IG .
2 G 2

(28)

Cylinder on a ramp

Let us consider a uniform cylinder of weight W and radius R rolling without slipping down a ramp of angle
.

We consider the xed reference frame, x, y, instantaneously located at the center of mass as shown in the
gure. The equations of motion, 22 and 23, are, in this case,
mx
G

= W sin F

myG

= N W cos

IG

F R

In these equations, W = mg, but the normal force, N , and the friction force, F , are unknown. These two
additional unknowns can be determined if we provide two additional kinematic conditions. First, we have
that yG = 0, from which we can determine N as
N = W cos .
Second, since the cylinder rolls without sliding, we have that x
G = R. Solving for x
G , we obtain
x
G =

g sin
,
1 + (IG /mR2 )

10

(29)

and F = (IG x
G )/R2 . For the uniform cylinder, we have that IG = mR2 /2 and x
G = (2g sin )/3. If instead
of having a uniform disc, we had a uniform ring with all the mass concentrated at the rim, then IG = mR2
and x
G = (g sin )/2. Also, if there was no friction, F would be zero, the cylinder would not rotate, and the
acceleration would be that of a sliding mass point x
G = g sin .
We now consider that the total cylinder mass is located as a mass point in the center of the cylinder. Then
the moment of inertia about G is zero. The cylinder will roll down the ramp but it requires no moment to
cause this motion.

Writing the equations about the center of mass we obtain the result that since the moment of inertia is zero,
the moment about G must be zero. Since both g and N are directed towards the center of mass, the only
moment comes from the friction F ; therefore, F must be zero.

Rotation about a Fixed Axis


For cases in which there is a xed point in the body, the motion of the body can be described with a single
parameter (e.g. the rotation angle). In principle, we could still consider equations (22) and (23) and use the
kinematic conditions to enforce that the motion of the xed point is zero. Alternatively, the analysis is often
simplied if we consider the conservation of angular momentum about the xed point directly. In this case,
we have,
IO = M O .

(30)

where IO is the moment of inertia about the xed point, O, is the angular acceleration, and M O is the
sum of external moments about O.

Example

Compound Pendulum

An example of a rigid body rotating about a xed axis is the compound pendulum. A compound pendulum
is a rigid body hinged, without friction, about a horizontal axis oset from its center of mass, and acted
upon by its own weight as an external force. The angular velocity of the pendulum about its pivot point is
= ; the angular acceleration of the pendulum about its pivot point is = . As noted before, every mass

11

point in the 2D solid body of the pendulum has the identical angular velocity and angular acceleration,
and .

It is convenient to apply the conservation on angular momentum about the point O because the unknown
reaction forces at the support contribute no moment about O. The conservation of angular momentum
about O results in
IO = MO = mgL sin ,
or,
g
sin .
IO /mL
Comparing it with the equation for a simple pendulum from Lecture 10, we see that the motion of a compound
=

pendulum is identical to the motion of a simple pendulum of equivalent length, Lequiv ,


Lequiv =

IO
.
mL

We now apply conservation of angular momentum about the center of mass to determine the reaction force
F . We see that gravity produces no moment about G and the only moments are produced by reaction forces
at the attachment point. The tension T at the attachment point produces no moment about the center
We remark on our earlier
of mass. The force F produces a moment MG = F L which must equal IG .
observation from Lecture 10 that if the pendulum is a mass point, the moment of inertia about the center
of mass is zero. Therefore, there can be no moment exerted by the attachment point about the center of
mass. Therefore can be no reaction in the direction normal to the direction from the pivot to the center
of masswhich would produce a moment. Therefore the force is aligned with the direction of the tension T
as in a string. We now apply Newtons law about the center of mass in the direction of T , balancing the
tension and a component of the gravitational force with the centripetal acceleration to determine T .
T + mg cos = mL2

(31)

ADDITIONAL READING
J.L. Meriam and L.G. Kraige, Engineering Mechanics, DYNAMICS, 5th Edition
4/1, 4/2, 4/3 (kinetic energy expression only), 4/4, 4/5 (momentum only) 6/1, 6/2, 6/3, 6/4, 6/5 (angular
momentum)

12

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

S. Widnall, J. Peraire
16.07 Dynamics
Fall 2008
Version 2.0

Lecture L12 - Work and Energy


So far we have used Newtons second law F = ma to establish the instantaneous relation between
the sum of the forces acting on a particle and the acceleration of that particle. Once the acceleration
is known, the velocity (or position) is obtained by integrating the expression of the acceleration
(or velocity). Newtons law and the conservation of momentum give us a vector description of the
motion of a particle in three dimensions
There are two situations in which the cumulative eects of unbalanced forces acting on a particle
are of interest to us. These involve:
a) forces acting along the trajectory. In this case, integration of the forces with respect to the
displacement leads to the principle of work and energy.
b) forces acting over a time interval. In this case, integration of the forces with respect to the
time leads to the principle of impulse and momentum as discussed in Lecture 9.
It turns out that in many situations, these integrations can be carried beforehand to produce
equations that relate the velocities at the initial and nal integration points. In this way, the
velocity can be obtained directly, thus making it unnecessary to solve for the acceleration. We shall
see that these integrated forms of the equations of motion are very useful in the practical solution
of dynamics problems.

Mechanical Work
Consider a force F acting on a particle that moves along a path. Let r be the position of the
particle measured relative to the origin O. The work done by the force F when the particle moves
an innitesimal amount dr is dened as
dW = F dr .

(1)

That is, the work done by the force F , over an innitesimal displacement dr, is the scalar product
of F and dr. It follows that the work is a scalar quantity. Using the denition of the scalar product,
we have that dW = F dr = F ds cos , where ds is the modulus of dr, and is the angle between
F and dr. Since dr is parallel to the tangent vector to the path, et , (i.e. dr = ds et ), we have that
F et = Ft . Thus,
dW = Ft ds ,

(2)

which implies that only the tangential component of the force does work.
During a nite increment in which the particle moves from position r 1 to position r 2 , the total
work done by F is

r2

s2

F dr =

W12 =
r1

Ft ds .

(3)

s1

Here, s1 and s2 are the path coordinates corresponding to r 1 and r 2 .


Note

Units of Work

In the international system, SI, the unit of work is the Joule (J). We have that 1 J = 1 N m. In
the English system the unit of work is the ft-lb. We note that the units of work and moment are
the same. It is customary to use ft-lb for work and lb-ft for moments to avoid confusion.

Principle of Work and Energy


We now a consider a particle moving along its path from point r 1 to point r 2 . The path coordinates
at points 1 and 2 are s1 and s2 , and the corresponding velocity magnitudes v1 and v2 .

If we start from (3) and use Newtons second law (F = ma) to express Ft = mat , we have

s2

W12 =

s2

Ft ds =
s1

v2

mat ds =
s1

v1

dv ds
m
ds =
ds dt

v2

v1

1
1
mv dv = mv22 mv12 .
2
2

(4)

Here, we have used the relationship at ds = v dv, which can be easily derived from at = v and v = s
(see lecture D4).
Dening the kinetic energy1 , as
1
T = mv 2 ,
2
we have that,
W12 = T2 T1

or

T1 + W12 = T2 .

(5)

The above relationship is known as the principle of work and energy, and states that the mechanical
work done on a particle is equal to the change in the kinetic energy of the particle.

External Forces
Since the body is rigid and the internal forces act in equal and opposite directions, only the external
forces applied to the rigid body are capable of doing any work. Thus, the total work done on the
body will be

i=1

(Wi )12 =

i=1

(r i )2

F i dr ,

(r i )1

where F i is the sum of all the external forces acting on particle i.


Work done by couples
If the sum of the external forces acting on the rigid body is zero, it is still possible to have non-zero
work. Consider, for instance, a moment M = F a acting on a rigid body. If the body undergoes
a pure translation, it is clear that all the points in the body experience the same displacement,
1

The use of T to denote the kinetic energy, instead of K, is customary in dynamics textbooks

and, hence, the total work done by a couple is zero. On the other hand, if the body experiences a
rotation d, then the work done by the couple is
dW = F

a
a
d + F d = F ad = M d .
2
2

If M is constant, the work is simply W12 = M (2 1 ). In other words, the couples do work
which results in the kinetic energy of rotation.

External and Internal Forces


In a typical dynamical system, the force F is composed of two terms: an external force F E and an
internal force F I . The external force results from an external actor applying an arbitrary forceof
his choice to the system. The external force does work and changes the energy of the system at
the whim of the actor. When the external force is removed, the system may oscillate or otherwise
move subject to internal forces. Internal forces are of two types: conservative internal forces such
as gravity which conserve energy but for example can transform kinetic into potential energy and
vice versa; and friction which acts internally to dissipate energy from the system. Initial conditions
are often set by applied external forces to the system, such as doing work by moving a pendulum
through an initial angle 0 . When they are removed, the system oscillates perhaps conserving
energy if friction is absent. We will consider conservative forces shortly.
Example

Block on an incline

Consider a block resting on an incline at position 1 in the presence of gravity. The gravitational
force acting in the vertical direction is mg. The block is supported on the plane by a normal
force FN = mgsin. We desire to push the block up to position 2. To do this, we must apply
an external force to overcome the component of gravity (internal force) along the plane plus the

friction force (internal force) acting to oppose the motion. In this process, the normal force does
no work.

The total tangential force acting on the block is then


FT = FE + FI

(6)

with
FE = mgsin mgcos

(7)

FI = +mgsin + mgcos

(8)

for a total force, internal plus external, of zero. Under these assumptions, the block arrives at
position 2 with zero velocity. Therefore, by equation (5), no work is done. At rst, this does not
agree with our intuition. We certainly felt we did work in pushing the block up the plane. But
the work done by the friction force and gravitational force exactly canceled this work. We also feel
that having raised the block to a higher position, there is some inherent gain in energy which
could be collected in the future. This is true! This result that the total force is zero resulting in
no work and no kinetic energy being gainedis due to the external application of force. These
external forces exactly canceled the internal forces of gravity and friction, driving the total force
to zero and resulting in no total work In other words, we became an actor instead of an observer.
These issues will be considered further when we consider conservative forces and potentials

Example

Block sliding down an incline

Having applied an external force FE to move the block to a higher position on the ramps where it
rests with no kinetic energy, we now become an observer and release it with only the internal forces
of gravity and friction acting. The coecient of kinetic friction between the surface of the ramp
and the block is . We want to determine the velocity of the block as a function of the distance
traveled on the ramp, s.

The forces on the block are: the weight, mg, the normal force, N , and, the friction force, N . We
have that Fn = man and Ft = mat . Since Fn = N mg cos and an = 0, we have N = mg cos .
Thus, Ft = mg sin N = mg sin mg cos , which is constant. If we apply the principle of
work and energy between the position (1), when the block is at rest at the top of the ramp, and
the position (2), when the block has traveled a distance s, we have T1 = 0, T2 = (mv 2 )/2, and the
work done by Ft is simply W12 = Ft s. Thus,
T1 + W12 = T2 ,

or,

1
mg(sin cos )s = mv 2 .
2

From which we obtain, for the velocity,


v=

2g(sin cos )s .

We make two observations: rst, the normal force, N , does no work since it is, at all times,
perpendicular to the path, and second, we have obtained the velocity of the block directly without
having to carry out any integrations. Note that an alternative, longer approach would have been
to directly use F = ma, and integrate the corresponding expression for the acceleration.

Rolling Cylinder, Friction Forces, Work


The cylinder rolling on a at plane is a very basic conguration in dynamics. As noted in Lecture
2, it is a single degree of freedom system with a denite relationship between the position of the
center of the cylinder, x0 (t) and the rotation angle (t): (t) = x0 (t)/R. The kinematics of the
rolling cylinder are shown in a). Consider a mass point at the edge of the cylinder. The dashed
curve shows the path taken by this mass point. Of signicance is the behavior near the plane. The
point O is an instantaneous center of rotation; the tangential velocity of the cylinder is zero about
this point. The acceleration of the mass point is not zero, nor is its vertical velocity.

The dynamics of the rolling cylinder is shown in b). If the cylinder moves with constant veloc
ity, nothing more need be said. However, if x
0 is not zero, then = , the angular acceleration
will be nonzero. This will require a moment about the center of mass, which in the simplest
conguration sketched, can only come from friction with the plane. Ff R = IG . Since the
point in contact with the plane is an instantaneous center of rotationdoes not movethis fric
tion force does no work. Also shown in b) are a variety of congurations of rolling cylinders.
The solid cylinder has a moment of inertia of IG = 1/2mR2 ; the cylinder whose mass is con
centrated in the rim, has a moment of inertia IG = mr2 ; the cylinder whose mass point is con
centrated in the center has a moment of inertia of zero. Therefore, no moment is required to
change its angular velocity and it behaves like a mass point moving on a frictionless surface. The
collision of two mass points can easily be realized by the collision of two such rolling cylinders.
Note

Cylinder rolling down a ramp

In parallel with our discussion of a block sliding down a ramp in the presence of friction, we now
consider a cylinder rolling down a ramp in the presence of friction. We assume that friction forces
are large enough to keep the kinematic relationship between the velocity of the cylinder and the
angular velocity of the cylinder intact. The cylinder is located at an initial position 1 as shown,
and is at rest. It is released and rolls down to position 2 where we observe it. The forces acting on
the cylinder are gravity, the normal force N, and the friction force Ff . Although the friction force
is necessary to keep the kinematic relationship intact, it does no work; as before, the normal
force does no work. Therefore, the only work is done by gravity. Thus we can write
T1 + W12 = T2

(9)

The initial kinetic energy is zero. The work done by gravity is mg sin s; the nal kinetic energy,
which includes both the kinetic energy due to translation and the kinetic energy due to rotation is
T2 = 1/2mv 2 + 1/2IG (v/R)2 . Therefore, the nal velocity is

2 g sin s
v=
1 + IG /R2

Note

(10)

Alternative expressions for dW

We have seen in expression (2) that a convenient set of coordinates to express dW are the tangentialnormal-binormal coordinates. Alternative expressions can be derived for other coordinate systems.
For instance, we can express dW = F dr in:
8

cartesian coordinates,

dW = Fx dx + Fy dy + Fz dz ,

cylindrical (polar) coordinates,


dW = Fr dr + F rd + Fz dz ,
or spherical coordinates,
dW = Fr dr + F r cos d + F rd .
As an illustration, lets calculate the work done by a constant internal force, such as that due to
gravity. The force on a particle of mass m is given by F = mgk. When the particle moves from
position r 1 = x1 i + y1 j + z1 k to position r 2 = x2 i + y2 j + z2 k, work is done, and the work may be
written as

r2

z2

F dr =

W12 =
r1

mg dz = mg(z2 z1 ) .
z1

Power
In many situations it is useful to consider the rate at which a device can deliver work. The work
per unit time is called the power, P . Thus,
P =

dW
dr
=F
=F v .
dt
dt

The unit of power in the SI system is the Watt (W). We have that 1 W = 1 J /s. In the English
system the unit of power is the ft-lb/s. A common unit of power is also the horse power (hp), which
is equivalent to 550 ft-lb/s, or 746 W.
Note

Eciency

The ratio of the power delivered out of a system, Pout , to the power delivered in to the system, Pin ,
is called the eciency, e, of the system.
e=

Pout
.
Pin

This denition assumes that the energy into and out of the system ows continuously and is not
retained within the system. The eciency of any real machine is always less than unity since there
is always some mechanical energy dissipated as heat due to friction forces.

ADDITIONAL READING
J. B. Marion, S. T. Thornton Classical Dynamics of Particles and Systems, Harcourt Brace,
New York, Section 2.5
J.L. Meriam and L.G. Kraige, Engineering Mechanics, DYNAMICS, 5th Edition

10

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

S. Widnall, J. Peraire
16.07 Dynamics
Fall 2008
Version 2.0

Lecture L13 - Conservative Internal Forces and Potential Energy


The forces internal to a system are of two types. Conservative forces, such as gravity; and dissipative forces

such as friction. Internal forces arise from the natural dynamics of the system in contract to external forces

which are imposed from an external source.

We have seen that the work done by a force F on a particle is given by dW = F dr.

If the work done by an internal forces F , when the particle moves from any position r 1 to any position r 2 ,
can be expressed as the dierence in a scalar function of r between the two ends of the trajectory,
r2
W12 =
F dr = (V (r 2 ) V (r 1 )) = V1 V2 ,

(1)

r1

then we say that the force is conservative. In the above expression, the scalar function V (r) is called the
potential. It is clear that the potential satises dV = F dr (the minus sign is included for convenience).
There are two main consequences that follow from the existence of a potential: i) the work done by a
conservative force between points r 1 and r 2 is independent of the path. This follows from (1) since W12
only depends on the initial and nal potentials V1 and V2 (and not on how we go from r 1 to r 2 ), and
ii) the work done by potential forces is recoverable. Consider the work done in going from point r 1 to
point r 2 , W12 . If we go, now, from point r 2 to r 1 , we have that W21 = W12 since the total work
W12 + W21 = (V1 V2 ) + (V2 V1 ) = 0.

In one dimension any force which is only a function of position is conservative. That is, if we have a force,
F (x), which is only a function of position, then F (x) dx is always a perfect dierential. This means that we
can dene a potential function as

V (x) =

F (x) dx ,
x0

where x0 is arbitrary.

In two and three dimensions, we would, in principle, expect that any force which depends only on position,

F (r), to be conservative. However, it turns out that, in general, this is not sucient. In multiple dimensions,

the condition for a force eld to be conservative is that it can be expressed as the gradient of a potential
function. That is,
F C = V .
This result follows from the gradient theorem, which is often called the fundamental theorem of calculus,
which states that the integral

r2

V dr = (V2 V1 )
r1

is independent of the path between r1 and r2 . Therefore the work done by conservative forces depends only
upon the endpoints r2 and r1 rather than the details of the path taken between them.
r2
r2
F C dr =
V dr = (V2 V1 )
r1

r1

In the general case, we will deal with internal forces that are a combination of conservative and nonconservative forces.
F = F C + F N C = V + F N C .

The gradient operator,

Note
The gradient operator, (called del), in cartesian coordinates is dened as
( )

( )
( )
( )
i+
j+
k.
x
y
z

When operating on a scalar function V (x, y, z), the result V is a vector, called the gradient of V . The
components of V are the derivatives of V along each of the coordinate directions,
V

V
V
V
i+
j+
k.
y
z
x

If we consider a particle moving due to conservative forces with potential energy V (x, y, z), as the particle
moves from point r = xi + yj + zk to point r + dr = (x + dx)i + (y + dy)j + (z + dz)k, the potential energy
changes by dV = V (x + dx, y + dy, z + dz) V (x, y, z). For small increments dx, dy and dz, and dV , can be
expressed, using Taylor series expansions, as
dV =

V
V
V
dx +
dy +
dz = V dr ,
x
y
z

where dr = dxi + dyj + dzk.


This equation expresses the fundamental property of the gradient. The gradient allows us to nd the change
in a function induced by a change in its variables.
If we write V (x, y, z) = C, for some constant C, this is the implicit equation of a surface, which is called a
constant energy surface. This surface is made up by all the points in the x, y, z space for which the function
V (x, y, z) is equal to C. It is clear that if a particle moves on a constant energy surface, dV = 0, since V is

constant on that surface. Therefore, when a particle moves on a constant energy surface, dr will be tangent
to that surface, and since
0 = dV = V dr ,
we have that V is perpendicular to any tangent to the surface. This situation is illustrated in the picture
below for the two dimensional case. Here, the constant energy surfaces are contour curves, and we can see
that the gradient vector is always normal to the contour curves.

Note

Gradient operator in cylindrical coordinates

The gradient operator can be expressed in cylindrical coordinates by writing x = r cos , y = r sin , and

r = x2 + y 2 , = tan1 (y/x). Thus, applying the chain rule for dierentiation, we have
( )
x
( )
y

=
=

r ( )
r ( )
( ) sin ( )

+
= cos
x r
x
r
r r
r ( ) r ( )
( ) cos ( )
+
= sin
+
.
y r
y
r
r r

If we note that i = cos er sin e and j = sin er + cos e , we have that


( )

( )
1 ( )
( )
er +
e +
.
r
r
z

An expression for spherical coordinates can be derived in a similar manner.

Conservation of Energy
When all the forces doing work are conservative, the work is given by (1), and the principle of work and
energy derived in the last lecture,
T1 + W12 = T2 ,
reduces to,
T1 + V1 = T2 + V2
3

or more generally, since the points r 1 and r 2 are arbitrary,


E = T + V = constant .

(2)

Whenever applicable, this equation states that the total energy stays constant, and that during the motion
only exchanges between kinetic and potential energy occur.
In the general case, however, we will have a combination of conservative, F C , and non-conservative, F N C ,
forces. In this case, the work done by the conservative forces will be calculated using the corresponding
C
= V1 V2 , and the work done by the non-conservative forces will be path
potential function, i.e., W12

dependent and will need to be calculated using the work integral. Thus, in the general case, we will have,
r2
T1 + V1 +
F N C dr = T2 + V2 .
r1

The work done by non-conservative forces which oppose the motion is negative. Therefore the sum of T2 + V2
will be less than T1 + V1 .

Examples of Conservative Forces


Gravity near the earths surface
On a at earth, the specic gravity g points down (along the -z axis), so F = mgk. Call V = 0 on the
surface z = 0, and then

V (z) =

(mg) dz,

V (z) = mgz .

For the motion of a projectile, the total energy is then


E=

1
mv 2 + mgz = constant .
2

Since vx and vy remain constant, we also have

1
mv 2 + mgz = constant.
2 z

Gravity
In a central gravity eld
F = G

Mm
Mm
er = (G
),
2
r
r

and so, taking V (r ) = 0,


V = G

Mm

= m.
r
r

where G is the universal gravitational constant and = M G is the strength of the gravitational eld from
a central body of mass M .

Spring Force
For small displacement, the force supported by a spring is F = kx. The elastic potential energy of the
spring is the work done on it to deform it an amount x. Thus, we have
x
1
V =
kx dx = kx2 .
2
0
If the deformation, either tensile or compressive, increases from x1 to x2 during the motion, then the change
in potential energy of the spring is the dierence between its nal and initial values, or,
V =

1
k(x22 x21 ) .
2

Gravity Potential for a Rigid Body


In this case, the potential Vi associated with particle i is simply Vi = mi gzi , where zi is the height of particle
i above some reference height. The force acting on particle i will then be F i = Vi . The work done on
the whole body will be
n

i=1

r 2i

r 1i

F i dr i =

((Vi )1 (Vi )2 ) =

i=1

mi g((zi )1 (zi )2 = V1 V2 ,

i=1

where the gravity potential for the rigid body is simply,


V =

mi gzi = mgzG ,

i=1

where zG is the z coordinate of the center of mass. Its obvious but worth noting that because the gravitational
potential is taken about the center of mass, the inertia plays no role in determining the gravitational potential.

Example

Cylinder on a Ramp

We consider a homogeneous cylinder released from rest at the top of a ramp of angle , and use conservation
of energy to derive an expression for the velocity of the cylinder.

Conservation of energy implies that T +V = Tinitial +Vinitial . Initially, the kinetic energy is zero, Tinitial = 0.
Thus, for a later time, the kinetic energy is given by
T = Vinitial V = mgs sin ,
5

where s is the distance traveled down the ramp. The kinetic energy is simply T =

1
2
2 IC ,

where IC =

IG + mR2 is the moment of inertia about the instantaneous center of rotation C, and is the angular
velocity. Thus, IC 2 = 2mgs sin , or,
v2 =

2gs sin
,
1 + (IG /mR2 )

since = v/R. For the general case of a cylinder with the center of mass at the center of the circle but an
2
uneven mass distribution, we write T = 12 m(1 + kG
/R2 ), where the eect of mass distribution is captured

in kG ; the smaller kG , the more concentrated the mass about the center of the cylinder. Then
v2 =

2gy
2 /R2
1 + kG

(3)

where s sin has been replaced with the vertical distance y. This equation shows that the more the mass
is concentrated towards the center of the cylinder (kG small), a higher velocity will be reached for a given
height, i.e less of the potential energy will go into rotational kinetic energy.

Example

Principle of Work and Energy

The m = 30 kg collar is released from rest at B and slides with negligible friction up the xed rod inclined
60o from the horizontal under the action of a constant force F = 450N applied to the cable. We want to
calculate the required stiness k of the spring so that its maximum deection equals 5cm. The position of
the small pulley at C is xed.

First, we want to calculate the work done by the cable. When the collar is at B, the length of cable between

the collar and the pulley is (12 + 0.22 ) = 1.0198 m. When the collar reaches its nal position, the length
of cable between the collar and the pulley is 0.2m. Since the force is constant, the work done by the force
F on the collar is simply Wcable = 450(1.0198 0.2) = 368.9118 Nm. Applying the principle of work and
energy between the initial position and the point of maximum spring compression (denoted by the subscript
f ), we have
TB + VB + Wcable = Tf + Vf .
Here, the kinetic energy at B and f is zero since both the initial and nal velocities are zero. We can
arbitrarily set the potential energy at B equal to 0. The potential energy at f will be due to gravity and to
6

the compression of the spring. Thus, we will have Vf = mg(1 + ) sin(60o ) + k 2 /2. Or,
1
368.9118 = (30)(1 + 0.05) sin(60o ) + k 0.052
2

k = 81038N/m

Note that when calculating the length of the cable in the nal state, we have neglected the compression of
the spring. This eect could easily be taken into account and the result would not dier much from the one
obtained here.

Equilibrium and Stability


If all the forces acting on the body are conservative, then the potential energy can be used very eectively
to determine the equilibrium positions of a system and the nature of the stability at these positions. Let us
assume that all the forces acting on the system can be derived from a potential energy function, V . It is
clear that if F = V = 0 for some position, this will be a point of equilibrium in the sense that if the
body is at rest (kinetic energy zero), then there will be no forces (and hence, no acceleration) to change the
equilibrium, since the resultant force F is zero.
Once equilibrium has been established, the stability of the equilibrium point can be determine by examining
the shape of the potential function. If the potential function has a minimum at the equilibrium point, then
the equilibrium will be stable. This means that if the potential energy is at a minimum, there is no potential
energy left that can be traded for kinetic energy. Analogously, if the potential energy is at a maximum, then
the equilibrium point is unstable.
Let us consider a particle under the eect of a potential force. The result F = V is useful not only for
computing the force but also for computing the stability of the motion from a diagram of the potential energy.
For instance, in the case of a particle attached at the end of a spring the potential energy is V = 12 kx2 .

At a point x > 0, V = dV /dx > 0 and so the force is negative. Similarly for x < 0 the force is positive.
At x = 0, dV /dx = 0 and the force is zero. We see that the force is directed towards the origin no matter
which way the particle is displaced and the force is only zero at the origin. The minimum of the potential
energy coincides with the equilibrium position of the particle. It is clearly a stable equilibrium, since any
displacement of the particle produces a force which tends to push the particle toward its resting point.

When V = 0 the system is in equilibrium. However, if this occurs at a maximum of V , the equilibrium is
not stable, since a positive displacement produces a positive force that tends to increase the displacement.
A pendulum of length L supporting as mass m is a good illustration of this.

If we take the potential energy to be zero at the bottom of its swing, we see that
V () = mgL(1 cos ) .
The pendulum is in equilibrium for = 0 and = . However, only = 0 is a stable minimum since
it is the only one that corresponds to a minimum.(Equilibrium requires that no force acts on the particle:
F () =

dV ()
d

= 0.) When the potential is a function of just one variable (e.g. x or ), there is a simple test

that can be used to determine if the equilibrium points (i.e. points where dV /dx = 0) are stable or unstable.
This test is based on looking at the value of the second derivative of the potential at the equilibrium point.
That is if d2 V /dx2 > 0 then the equilibrium point corresponds to a minimum of the potential energy and
therefore, the equilibrium is stable. When d2 V /dx2 < 0 then the equilibrium point occurs at a maximum of
the potential function and the equilibrium point is unstable. The test only breaks down when d2 V /dx2 = 0.
In this case, we would need to look at higher derivatives to determine the stability of the system.
Example

Equilibrium and Stability

A cylinder of radius R, for which the center of gravity, G, is at a distance d from the geometric center, C,
lies on a rough plane inclined at an angle .

Since gravity is the only external force acting on the cylinder that is capable of doing any work, we can
examine the equilibrium and stability of the system by considering the potential energy function. We have

zC = zC0 R sin , where zC0 is the value of zC when = 0. Thus, since d = |CG|, we have,
V = mgzG = mg(zC + d sin ) = mg(zC0 R sin + d sin ) .
The equilibrium points are given by V = 0, but, in this case, since the position of the system is uniquely
determined by a single coordinate, e.g. , we can write
V =

dV
,
d

which implies that, for equilibrium, dV /d = mg(R sin + d cos ) = 0, or, cos = (R sin )/d. If
d < R sin , there will be no equilibrium positions. On the other hand, if d R sin , then eq. =
cos1 [(R sin )/d] is an equilibrium point. We note that if eq. is an equilibrium point, then eq. is also an
equilibrium point (i.e. cos = cos()).

In order to study the stability of the equilibrium points, we need to determine whether the potential energy
is a maximum or a minimum at these points. Since d2 V /d2 = mgd sin , we have that when eq. < 0,
then d2 V /d2 > 0 and the potential energy is a minimum at that point. Consequently, for eq. < 0, the
equilibrium is stable. On the other hand, for eq. > 0, the equilibrium point is unstable.
Example

Tipped Cylinder and Ellipse

Consider the solid semi-circle at rest on a at plane in the presence of gravity. At rest, it is in equilibrium
since the gravitational moments balance. We consider that it tips and rolls, keeping the no-slip condition
satised. To determine the stability, we consider the change in potential energy, V (). Only the vertical
displacement of the center of mass contributes to a change in potential. If we expand the potential V ()
for small , we will get an expression V () = A2 . (Recall that for the pendulum, V () = mgL2 /2.) The
question of stability depends upon the sign of A. If A is positive, the system is stable; if A is negative, the
system is unstable.
When the cylinder tips, this motion results in a vertical displacement of the center of mass, y and a
horizontal displacement of the center of mass x, where y and x can be found from the geometry.
Consider the case where the center of mass is a distance L from the center of rotation of the cylinder.
Then, from the gure, we see that the cylinder rolls so that the point of contact is now at x = R0 . Then
x = R0 L sin 0 and y = L(1 cos 0 ). The vertical displacement of the center of mass is similar
to that of a pendulum of length L. The tipped cylinder is stable. If the center of mass is at the center of
rotation, r = 0, all angles 0 are points of neutral stability.
9

We now consider the two systems shown in the gure. These are simply semi-ellipses resting on a at plane.

Again, the point of symmetry will be an equilibrium point since the gravitational moments will balance.
But the question of stability relates to whether the center of mass moves up or down as increases. We
have V () = A2 , with stability for A > 0 and instability for < 0. We feel instinctively, that one of these
systemsthe tall skinny one might be unstable. This implies that it will not remain balanced about the
equilibrium point, but will tip over. For the cylinder, the radius R played an important role. It was the
distance from the point of contact to the center of curvature of the cylinder at the contact point. In this
more complex example, the role of the radius R is played by the radius of curvature at the contact point .
Since to determine stability, we consider only small displacements the curve may be considered as a local
cylinder. Referring to the gure, we see that the motion of the center of mass due to tipping of the ellipse
depends on the relation between the local radius of curvature and the distance of the center of mass from
the center of curvature, the center of rotation. If the center of mass lies below the center of curvature, the
small displacement motion will be stable, much like a pendulum. If the center of mass lies about the center
of curvature, the motion will be unstable and the ellipse will initially tip over. From the gure, we see that
the radius of curvature is largest for the at ellipse and smallest for the tall ellipse, agreeing with our
intuition about which one is more likely to tip over. However, if the center of mass of the tall ellipse is below
10

its center of curvature/rotation, the ellipse will be stable.

Energy Diagrams(KK)
Energy diagrams provide a useful way to study the motion of conservative one dimensional systems. In a
conservative system, the total energy E is a constant; the motion transforms the form of the energy from
kinetic to potential while keeping the total constant. For a given position of the system, x, the potential
potential energy can be plotted, V (x). The total energy of the system is constant, and is also shown in the
diagram. Since the sum of the kinetic energy and the potential energy is a constant as the system moves
in x, the kinetic energy T = E V is easily found by inspection. Since the kinetic energy can never be
negative, the motion is constrained to regions where V E.

Since the system is conservative, the total energy E is constant. The kinetic energy T is greatest at the
origin x = 0. As the particle goes past the origin in either direction, it is slowed by the spring and comes to
a complete rest at one of the turning points x0 . The particle then moves to the origin increasing its kinetic
energy, and the cycle is repeated. We see that in the case of a harmonic oscillator the motion is always
bounded. As E increases the turning points move farther and farther o, but the particle remains bounded.
Also, note that when E = 0 then the particle is at x = 0 and the particle lies at rest in equilibrium.
Example

van der Waals Force

The situation is dierent when the function V does not increase indenitely with distance. Consider for
instance the interaction between two atoms. At large separations the atoms attract each other weakly with
the van der Waals force, which varies as 1/r7 . As the atoms approach the electron clouds begin to overlap
creating strong repulsive forces. The corresponding potential is given by

r 6
r0 12
0
V (r) =
2
r
r

11

For a positive energy E > 0, the motion is unbounded and the atoms are free to y apart. As the diagram
shows, the distance of closest approach ra does not change appreciably as E is increased. The situation is
quite dierent for E < 0. In this case the motion is bounded for small and large separations. The atoms
never approach closer then rb and they never move apart farther than rc . A bound system of two atoms
is a molecule. If two atoms collide with positive energy they cannot form a molecule unless some means is
available for losing energy to make E negative. In general a third body is necessary to carry o the excess
energy.

Small Oscillations in a bound system (KK)


Every bound system oscillates as a harmonic oscillator about its equilibrium position if it is perturbed from
the equilibrium position by a small amount. This can be seen by noting that the minimum of the potential
energy can be generally approximated by a parabola in the neighborhood of the minimum.

If the total energy is low enough so that the motion is restricted to the region where the curve is nearly
parabolic, the system will behave like a harmonic oscillator. If V (r) is well behaved and has a minimum at
12

r0 , then we can always expand it in Taylors series about point r0 . Thus,

2
dV
1
2 d V
V (r) = V (r0 ) + (r r0 )
+
(r

r
)
0
dr r0 2
dr2

+ ...
r0

However, since at r0 , dV /dr = 0, for suciently small displacements we can truncate the series and obtain,

2
1
2 d V
V (r) = V (r0 ) + (r r0 )
.
2
dr2 r0
k(r r0 )2
.
2

We can also identify the eective spring constant as k = d2 V /dr2 r0 . These ideas can be applied to many
V (r) = constant +

systems, identifying the oscillatory behavior by considering the behavior of the potential function near the
equilibrium point, the minimum of the potential function. (The term eective is used to emphasize that
the stiness in a system can be due to many eects: a spring, gravity, elasticity or a combination of these
eects.) The value of the constant plays no role in the dynamics of the system.

Small Displacements of a Mass-spring System


We now consider the potential for the familiar mass-spring system , previously discussed.

For small displacements of a mass spring system, whose equilibrium position is x = 0, the potential function
can be written

V (r) =

kx2
.
2

where k is the spring constant. In a harmonic oscillator without damping, such as the examples discussed
here, energy is conserved. As potential energy increases, kinetic energy decreases. Thus the minimum of
V occurs at the maximum of T. For small amplitude motions about x = 0, both the displacement x and
the velocity are sinusoidal in time: x(t) = Asin(t + 0 ), where A and 0 are determined from the initial
conditions. The kinetic energy is then given by
T =

1
1
mv(t)2 = m 2 (Acos(t + 0 ))2
2
2

(4)

where is the natural frequency of oscillation, that frequency which occurs as an unforced interchange
between kinetic and potential energy. The oscillation occurs symmetrically about x = 0 the minimum of the
13

potential function. Since the total energy E remains constant during the oscillation
k(Asin(t + 0 ))2
m 2 (Acos(t + 0 ))2
+
,
2
2

E = V (x(t)) + T (x(t)) =
and we obtain the result =

(5)

k
m.

We also observe that in this case the maximum value of V (x) is VM AX =

kA2
2 ;

this occurs at x = 0. The

m 2 A2
;
2

maximum value of T (x) is TM AX =


this occurs at x = A. Equating these maximum values we again

k
obtain =
m . Any constant added to V (x) plays no role in the dynamics. If V (0) = C, then VM AX
would be written VM AX = (V (x) V (0))M AX , thus removing the constant from consideration.

Pendulum
Earlier we derived the potential function for the pendulum as a function of the angle . We saw that the
pendulum exhibited a range of behavior from stable oscillation about = 0 to unstable divergence if the
initial position was near = . The potential function for the pendulum was given by
V () = mgL(1 cos ) .
Expanding cos about theta = 0 and keeping only the terms to order 2 for small displacements, we obtain
V () = mgL(1 cos ) = mgL(1 (1

2
2
+ .....)) = mgL
.
2
2

For small amplitude motions about = 0, both the displacement and the velocity are sinusoidal in time.

As potential energy increases, kinetic energy decreases. Thus the minimum of V occurs at the maximum of

T. The total energy is constant. The maximum value of kinetic energy is given by
2
1
d
1
2
TM AX = mL
= mL2 2 02
2
dt M AX
2

(6)

where is the natural frequency of oscillation and 0 is the maximum amplitude. The oscillation occurs
symmetrically about = 0, the minimum of the potential function. The maximum value of the potential is
VM AX = mgL

02
2 ;

this occurs at = 0 the point of maximum amplitude where the velocity is zero. Equating

the maximum value of kinetic energy to the maximum value of the potential
TM AX =

1
2
mL2 2 02 = VM AX = mgL 0
2
2
14

(7)

we obtain the result =

L.

Conversely if we expand the potential for small displacement near = , substituting = + , and
expanding V () for small we obtain
V () = mgL(1 cos( + )) = mgL(1 + cos()) = mgL(1 + (1

2
2
+ ....)) = mgL(2
)
2
2

(8)

()
This result completely changes the dynamics of the pendulum system. Since F = dVd
= mgL, a

positive force in the direction of motion would exist. This is equivalent to a negative spring. If there was
no additional restoring force, say from a spring opposing the pendulum motion, an unstable divergence
would occur. (Recall that any constant adding to the value of V () is of no signicance in determining the
dynamics of a system; only the slope and higher order derivatives play a role.)

Oscillating Tipped Cylinder/Ellipse


We now consider how to determine the frequency of oscillation of the tipped ellipse, the cylinder being just
a special case. It is obvious that the semi-circle will oscillate about its center of symmetry. The change in
the gravitational potential is given by the pendulum formula relating the position of the center of mass
L to the radius R, or for the more general case the formula relating position of the center of mass L to the
radius of curvature .
To determine the frequency, we need to identify TM AX , the maximum value of kinetic energy. Although
the moment of inertia played no role in determining the stability of the cylinder to tipping displacement,
the moment of inertia will contribute to kinetic energy and thus aect the frequency of oscillation. The
system has both translation and rotational kinetic energy, and both will be at their maximum values when
the system moves through the point of symmetry, = 0.
The maximum kinetic energy will be the sum of the translational and rotational kinetic energies. The
maximum translational kinetic energy will be the product of the maximum velocity of the center of mass
and the total mass of the cylinder; the maximum rotational kinetic energy will be the product of maximum
value of the angular velocity and the moment of inertia about the center of mass. Both will reach their
maximum when the system moves through = 0, the point of minimum gravitational potential.

References
[1] M. Martinez-Sanchez, Unied Engineering Notes, Course 95-96.
[2] D. Kleppner and R.J. Kolenkow, An Introduction to Mechancis, McGraw-Hill, 1973.
ADDITIONAL READING
J.L. Meriam and L.G. Kraige, Engineering Mechanics, DYNAMICS, 5th Edition
3/7

15

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

J. Peraire, S. Widnall
16.07 Dynamics
Fall 2008
Version 2.0

Lecture L14 - Variable Mass Systems: The Rocket Equation


In this lecture, we consider the problem in which the mass of the body changes during the motion, that is,
m is a function of t, i.e. m(t). Although there are many cases for which this particular model is applicable,
one of obvious importance to us are rockets. We shall see that a signicant fraction of the mass of a rocket
is the fuel, which is expelled during ight at a high velocity and thus, provides the propulsive force for the
rocket.
As a simple model for this process, consider the cases sketched in a) through f) of the gure. In a) we
consider 2 children standing on a stationary at car. At t = 0, they both jump o with velocity relative to
the at car of u. (While this may not be a good model for children jumping o a atcar, it is a good model
for the expulsion of mass from a rocket where the mass ow from the choked nozzle occurs at a relative
velocity u from the moving rocket.) In b) we consider that they jump o in sequence, each jumping with a
velocity u relative to the then velocity of the atcar, which will be dierent for the 2nd jumper since the car
has begun to move as a result of jumper 1. For a), we have a nal velocity of the at car as
2mu
M + 2m

(1)

mu
mu
+
M + 2m M + m

(2)

V2R =
for b) we have
V22 =

where M is the mass of the atcar, and m is the mass of the jumper. We can see the case b) gives a higher
nal velocity. We now consider cases with more blocks; we introduce the notation VN N to be the nal
velocity with N blocks coming o one at a time. We use the notation VN R to denote the reference velocity
if the N blocks come o together. If we consider 3 blocks, we can see by inspection that
V 3R =

3mu
M + 3m

(3)

and
mu
mu
mu
+
+
M + 3m M + 2m M + m

(4)

4mu
M + 4m

(5)

mu
mu
mu
mu
+
+
+
M + 4m M + 3m M + 2m M + m

(6)

V33 =
While for 4 blocks

V4R =
and
V44 =

Generalizing this to arbitrary N, we can write


VN R =
and
VN N =

N mu
M + Nm

i=1

mu
M + im

(7)

(8)

The gure shows the results for V100,100 /V100R , the increase in velocity as each block comes o. We choose
100 blocks with a mass m equal to .01 M; therefore the 100 blocks with m = .01 equal the mass M, the
propellant ratio of the rocket is 50%. The nal velocity of the cart is 40% higher than would be obtained
if the masses come o together. Momentum is conserved. The sum of the momentum of the cart plus all the
individual mass points is equal to zero. However, the momentum of all the particles left behind is of very
little interest to us. We care only about the nal velocity of the cart.
2

To analyze this question we must consider a system of variable mass, and the process by which it gains
velocity as a result of ejecting mass. While our previous discussion provided a discrete model of this process,
we now consider a continuous process, obtaining a relation for v(t) as a function of the momentum interaction
between the system and the external world.
We shall start by considering a body with velocity v and external forces F , gaining mass at a rate m
= dm/dt.
Let us look at the process of gaining a small amount of mass dm. Let v be the velocity of dm before it is
captured by m, and let f represent the average value of the impulsive forces that dm exerts on m during
the short interval dt, in which the capturing takes place. By Newtons third law, dm will experience a force
f , exerted by m, over the same dt.

We can now examine the capture process from the point of view of dm and equate the impulse, f dt, to
the change in linear momentum of dm,
f dt = dm(v + dv v ).

(9)

Here, v + dv is the velocity of m (and dm) after impact. Analogously, from the point of view of m, we write

F dt + f dt = m(v + dv) mv = mdv.

(10)

The term dm dv in equation (9) is a higher order term and will disappear when we take limits. The impulse
due to the contact force can be eliminated by combining equations (9) and (10),
F dt dm(v v ) = mdv,
or, dividing through by dt,

dv
dm
dm
= F (v v )
= F + (v v)
.
dt
dt
dt

(11)

Here, v v is the velocity of dm relative to m. This expression is valid when dm/dt > 0 (mass gain) and
when dm/dt < 0 (mass loss). If we compare this expression to the more familiar form of Newtons law for

a particle of xed mass m dv


dt = F , we see that the term (v v)dm/dt is an additional force on m which is

due to the gain (or loss) of mass. Equation (11) can also be written as
d(mv)
dm
= F + v
,
dt
dt
3

where v is the velocity of the captured (or expelled) mass relative to the velocity of the mass m. This shows
that, for systems involving variable mass, the usual expression stating conservation of linear momentum,
d(mv)/dt = F , is only applicable when the initial (nal) velocity of the captured (expelled) mass, v , is
zero. The behavior of m(t) is not an unknown, but is specied according to the characteristics of the rocket.
In most cases,

dm
dt

is a constant and negative. In some cases, the behavior of m(t) may be determined by a

control system. In any case, it is a given quantity.

The Rocket Equation


We consider a rocket of mass m, moving at velocity v and subject to external forces F (typically gravity
and drag). The rocket mass changes at a rate m
= dm/dt, with a velocity vector c relative to the rocket.
We shall assume that the magnitude of c is constant.

The velocity of the gas observed from a stationary frame will be v = v + c. In this frame, c is a vector
aligned along the ight path in a negative direction. c = ct, where t is the unit direction along the ight
path. Thus,

dv
dm
=F +c
.
dt
dt

(12)

The term T = c dm/dt is called the thrust of the rocket and can be interpreted as an additional force on the

rocket due to the gas expulsion.

Equation (12) is a vector equation which can be projected along the direction of v (tangent to the path).

Thus,

dv
dm
= Ft c
= Ft + T,
dt
dt

(13)

where Ft is the tangential component of F , v and c are the magnitudes of v and c respectively, and we have
assumed that c is parallel and has opposite direction to v. The magnitude of the thrust is T = cm
. Note
that for a rocket, m
will be negative (mass is lost). If the force Ft is known, this equation can be integrated
in time to yield an expression for the velocity as a function of time.
Let us consider some simple cases:

No External Forces: Ft = 0
If gravity and drag eects are neglected, we have,

dv
dm
= c
dt
dt

or, integrating between an initial time t0 , and a nal time t,

v = v v0 = c(ln mf ln m0 ) = c ln

m
m0
= c ln
.
mf
m0

(14)

Alternatively, this expression can be cast as the well known rocket equation,
m = m0 ev/c ,

(15)

which gives the mass of the rocket at a time t, as a function of the initial mass m0 , v, and c. The mass of
the propellant, mpropellant , is given by,
mpropellant = m0 m = m0 (1 ev/c ).
From the above equations, we see that for a given v and m0 , increasing c increases m (payload plus
structure) and decreases mpropellant . Unfortunately, we can only choose c as high as the current technology
will allow. For current chemical rockets, c ranges from 2500-4500 m/sec. Ion engines can have cs of roughly
105 m/sec. However, since relatively few particles are involved, the thrust is quite low, whereas chemical
rockets have high mass ows.
Equation (13) has some quite remarkable properties, as sketched below. First, the V for a given m0 /mf is
independent of time. More strikingly, if we compare the nal V to the at car situation where all the mass
was ejected instantaneously rather than continuously, we see a substantial gain in V from the continuous
process. However, the nal V for the continuous process is independent of the actual time span of the
process, as shown in the sketch. Thus in the limit as the continuous process occurs quickly, we will obtain
the result for the continuous process. We will refer to the rapid continuous process as a V impulse. When
we consider rocket thrusting during orbital maneuvers, we will consider only impulsive thrust.
A second property, is that the nal V from a series of rings is independent of the details and a function only
0
of the total mass change over the processes. Consider a rst thrust producing a V . Then V1 = c ln m
m1 .
1
Now the mass is m1 and a second thrust produces a V2 = c ln m
m2 . By a property of logarithms, adding

m1
m0
0
these two V s together gives Vf = V1 + V2 = c ln m
m1 + c ln m2 = c ln m2 where m2 is the nal mass

of the vehicle mf .

Another property is that the relationship between thrust and V is a positive scaler, always additive.
Whether we thrust in the positive or negative direction,or to the side it always consumes propellant, we
never get anything back
and substitute into equation 14 to obtain an
Note that if m
is constant, we can write m(t) = m0 + mt,
expression for v as a function of t,
v = v0 c ln(1 +

t)
m0

Recall that according to the convention used, m


is negative as the mass decreases with time.
Gravity: Ft = mg
A constant gravitational eld acting in the opposite direction to the velocity vector can be easily incorporated.
In this case, equation 13 becomes,
m

dv
dm
= mg c
,
dt
dt

which can be integrated to give


v = v0 c ln

m
m
m m0
gt = v0 c ln
g
.
m0
m0
m

(16)

This solution assumes that cm


> m0 g at t=0. If this is not true, the rocket will sit on the pad, burning fuel
until the remaining mass satises this requirement.
Gravity Plus Drag: Ft = mg D
The eect of the drag force, D, is harder to quantify. It turns out that for many important applications drag
eects are very small. The drag force is characterized in terms of a drag coecient, CD . Thus,
D=

1 2
v ACD ,
2
6

where A is the cross-sectional area of the rocket. The air density changes with altitude z, and may be
approximated by
= 0 ez/H ,
where H 8000 m is the so-called scale height of the atmosphere, and 0 is the air density at sea level.
It turns out that the dierential equation that results for the velocity cannot be integrated explicitly and,
in practice, needs to be integrated numerically. It is interesting to note, however, that the eect of drag
losses is usually quite small, and it is often reasonable to ignore it in a rst calculation. In order to see the
importance of D versus the eect of gravity, we can estimate the value of the ratio D/mg. At conditions
typical for maximum drag, 0.25 kg/m3 and v = 700 m/s. Considering a rocket of 12, 000 kg with a cross
section of A = 1 m2 and CD = 0.2, we have,
ACD v 2
0.25 1 0.2 7002
=
= 0.021 ,
2mg
2 12, 000 9.8
which indicates that the drag force is only about 2% of the gravity force.
Note (Optional)

Gravity Losses [1]

Let us consider a rocket providing a constant thrust T = cm


which is launched vertically upwards from
rest. Neglecting drag but considering gravity forces, the velocity is given by expression (16) which can be
re-written as

1
1
v = c ln
,

(17)

where we have introduced = m/m0 and the thrust induced acceleration measured in gees, n =
cm/(gm

0 ). Now, v = dz/dt, and the above equation can be integrated to give

c2
1
(1 )2
z=
1 ln
.
gn

2n

(18)

If we consider equations (17) and (18) at burnout time (when all the propellant is consumed), the mass
fraction will be = (m0 mpropellant )/m0 . After this time, the rocket will coast in free-ight. The total
energy per unit mass at this (and all subsequent times) is

2
v2
1
1
1 ln(1/)
2
E=
+ gz = c
ln
+
.
2
2

n
Since, 0 < < 1, 1 ln(1/) is negative and consequently, higher n means larger nal energy per unit
mass E, tending to the ideal limit
Eideal =

1
2

2
1
c ln
,

which is that obtained for an impulsive start. Reducing n means applying smaller thrust for a longer time,
and, as the above equation shows there is a price to be paid in the eventual energy of the payload. One way
to quantify this loss is to dene an equivalent impulsive velocity (v0 )eq such as to achieve the energy per

unit mass E in a very short time, hence, at ground level,

2
1
1 ln(1/)
(v0 )eq = c
ln
+2
,

n
and to compare this to the ideal impulsive velocity
(v0 )ideal = c ln

1
.

Some representative fractional losses ((v0 )ideal (v0 )eq )/(v0 )ideal , are shown below,
= 0.7

= 0.5

= 0.3

n = 1.5

0.3628

0.3188

0.2676

n=3

0.1615

0.1444

0.1235

n = 10

0.0456

0.0410

0.0354

The table shows losses of 15% for n = 3, 4% for n = 10. (Recall that n must be greater than 1 for
immediate lift o from the launch pad.) The losses are severe for n as low as 1.5. The gure shows a graph
of these losses as a function on n for = .8, .7, .5, and .3.

Example

Single vs. Two Stage Rockets [1]

Single stage
To achieve an orbital speed of v = 7600 m/s, we require an ideal v of about 9000 m/s where the extra
velocity is needed to overcome gravity and drag. Chemical rockets produce exhaust jets at velocities of
c 2500 4500 m/s. Using the higher c, if we wish to place a payload in orbit with a single stage rocket,
we have a mass ratio (mass at burn out, mf , divided by initial mass, m0 ) of
mf
= e9000/4500 = e2 = 0.135 .
m0
8

The mass mf must include all components of the rocket infrastructure, including the engine, empty tank,
guidance equipment, etc., as well as the payload. The mass of the propellant will be m0 mf = (1
0.135)m0 = 0.865m0 . If we assume that the tank plus engine are the main contributors to the weight of the
rocket and weigh 10% of the propellant, 0.087m0 , we have,
m1stage
payload = (0.135 0.087)m0 = 0.048m0 .
Clearly, there is not much margin here, and, in fact, single-stage-to-orbit vehicles have yet to be engineered
successfully. The alternative solution is to subdivide the rocket, so that empty tanks are dropped when they
are no longer needed.
Two stages
Consider now a two stage vehicle. We shall again assume that each empty tank plus its engine weighs 10%

of the propellant it carries. The required v is now subdivided into two vs of 4500 m/s each.

First stage

If m0 is the initial mass and m1 is the mass after burn out, we have,

m1 = e4500/4500 m0 = e1 m0 = 0.368m0
The fuel burnt will be m0 m1 = (1 0.368)m0 = 0.632m0 , and the weight of the tank and engine to be
dropped will be 0.0632m0 , leaving an initial mass for the second stage, m2 , of
m2 = 0.368m0 0.063m0 = 0.295m0 .
Second stage
After burn out,
mf = e1 m2 = e1 0.295m0 = 0.109m0 .
The fuel burnt in this stage will be m2 mf = (0.295 0.109)m0 = 0.186m0 and the weight of the tank
plus engine will be 0.019m0 , leaving for the payload
m2stage
payload = (0.109 0.019)m0 = 0.090m0 ,
which is still very small but about twice the size of the payload obtained for the single stage rocket.

Note

Launch Strategy

The earth rotates towards the East with a period of about 23.9 hours. This results in a tangential velocity
of vt = 466m/sec. at the equator. The corresponding value for higher latitudes is vt = 466m/sec Cos0 ,
where 0 is the latitude of position on the earth. In most situations, we desire to use this tangential velocity
to add to the energy available to place a satellite in orbit. Therefore, most launch sites are located as close

to the equator as possible and the launch occurs with a velocity component directed towards the east. The
table shows the latitude of the current major international launch sites. Of interest is Sea Launch, which
leaves the US west coast with a sea-going launch platform and control ship and launches very close to the
equator. Other criterion for launch site is safety; a failed launch should not put populations in danger.
28.5o

Kennedy

Tanegashima

30.4

Vandenburg

34.4

Xichang

28.25

Baikonur

45.6o

Kourou

5.2

Sea Launch

0 19 min

Note

Gravity Turn

The model we have introduced for rocket launch is incomplete if our goal is to place a spacecraft into orbit.
Orbital launch requires that the ight end with a roughly horizontal velocity at orbital speed. One useful
maneuver to accomplish this transition is called the gravity turn. In this maneuver, gravity acts to turn the
trajectory of the rocket towards the horizontal.

This maneuver oers two main advantages over a conventional thrust-controlled trajectory where the rockets
own thrust is used to steer the vehicle. First, any thrust used in changing the rockets direction is not being
used to accelerate the vehicle into orbit, constituting a loss which can be reduced by using gravity to steer
the vehicle onto its desired trajectory. Second, and more importantly, because the force of gravity is doing
the steering during the initial ascent phase of the launch the ship can maintain low or even zero angle of
attack. This minimizes transverse stress on the launch vehicle; allowing for a weaker, and thus lighter,
launch vehicle. For a constant thrust, and a variable mass, the details of the trajectory and motion require
a numerical solution. Using local intrinsic coordinate, the governing equations for the gravity turn are
v = T /m(t) g cos

(19)

v = g sin

(20)

10

Note

Specic Impulse

The specic impulse Isp , with units of seconds, is often used in practice to characterize the performance of a
rocket engine. The denition of the specic impulse is the magnitude of the thrust divided by the propellant
weight ow rate,
Isp =

T
c
= .
mg

Typical values of Isp are around 300 s for solid propellants and up to 500 s for higher energy fuels.

ADDITIONAL READING
J.L. Meriam and L.G. Kraige, Engineering Mechanics, DYNAMICS, 5th Edition
4/6, 4/7

References
[1] M. Martinez-Sanchez, Unied Engineering Notes, Course 95-96.

11

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

J. Peraire, S. Widnall
16.07 Dynamics
Fall 2008
Version 1.2

Lecture L15 - Central Force Motion: Keplers Laws

When the only force acting on a particle is always directed to


wards a xed point, the motion is called central force motion.
This type of motion is particularly relevant when studying the
orbital movement of planets and satellites. The laws which gov
ern this motion were rst postulated by Kepler and deduced from
observation. In this lecture, we will see that these laws are a con
sequence of Newtons second law. An understanding of central
force motion is necessary for the design of satellites and space
vehicles.

Keplers Problem
We consider the motion of a particle of mass m, in an inertial reference frame, under the inuence of a force,
F , directed towards the origin.

We will be particularly interested in the case when the force is inversely proportional to the square of the
distance between the particle and the origin, such as the gravitational force. In this case,
F =

mer ,
r2

where is the gravitational parameter, r is the modulus of the position vector, r, and er = r/r.

It can be shown that, in general, Keplers problem is equivalent to the two-body problem, in which two

masses, M and m, move solely due to the inuence of their mutual gravitational attraction. This equivalence

is obvious when M m, since, in this case, the center of mass of the system can be taken to be at M .

However, even in the more general case when the two masses are of similar size, we shall show that the
problem can be reduced to a Kepler problem.
Although most problems in celestial mechanics involve more than two bodies, many problems of practical
interest can be accurately solved by just looking at two bodies at a time. When more than two bodies are
involved, the problem is considerably more complicated, and, in this case, no general solutions are known.
The two body problem was studied by Kepler (1571-1630) who lived before Newton was born. His interest
was in describing the motion of planets around the sun. He postulated the following laws:
1.- The orbits of the planets are ellipses with the Sun at one focus
2.- The line joining a planet to the Sun sweeps out equal areas in equal intervals of time
3.- The square of the period of a planet is proportional to the cube of the major axis of its elliptical orbit
In this lecture, we will start from Newtons laws and verify that the above three laws can indeed be derived
from Newtonian mechanics.

Equivalence between the two-body problem and Keplers problem


Here we consider the problem of two isolated bodies of masses M and m which interact though gravitational
attraction. Let r M and r m denote the position vectors of the two bodies relative to a xed origin O. Since
the only force acting on the bodies is the force of mutual gravitational attraction, the motion is governed by
Newtons law with an equal and opposite force acting on each body.

Mm
er ,
r2
Mm
= G 2 er ,
r

M r M = G
mr m

(1)
(2)

where r = |r|, er = r/r, and G is the gravitational constant.

The position of the center of gravity, G, of the two bodies will be


rG =

M r M + mr m
.
M +m

(3)

Since the two bodies are isolated, we will have, from momentum conservation, that r G =constant, and
r G = 0. Therefore, the position of the center of gravity, at all times, can be found trivially from the initial
conditions.
If the position vector of m as observed by M , r = r m r M , is known, then the position vectors of M and
m could be computed as
rM = rG

m
r,
M +m

rm = rG +

M
r.
M +m

(4)

Therefore, since we know the position of the center of mass rG for all time, we shall show that the problem
of determining r M and r m is equivalent to that of determining r, the vector distance between them.
The governing equations for rm and rM are given in equation (1) and (2). Subtracting these two expressions,
we obtain,
r = r m r M = G

M +m
er ,
r2

(5)

or,
Mm
Mm
r = G 2 er .
M +m
r

(6)

The above expression shows that the motion of m relative to M is in fact a Kepler problem in which the
force is given by GM mer /r2 (this is indeed the real force), but the mass of the orbiting body (m in this
case), has been replaced by the reduced mass, M m/(M + m). Note that when M >> m, the reduced mass
becomes m. However, the above expression is general and applies to general masses M and m.
Alternatively, the above expression can be written as
mr = G

(M + m)m
er ,
r2

(7)

which is again a Kepler problem for an orbiting body of mass m, in which the gravitational parameter is
given by = G(M + m).

Example

Solution to the Two Body Problem

There are two approaches to the solution of the two-body problem. One is a direct numerical attack on
equations (1) and (2); the other is to use the analytic solution of the Kepler problem, equation(7), and
having found r(t), to use the equation for the position of the center of mass, r G (t) and equation (4) to
determine r m (t) and r M (t). The position of the center of mass is determined by the initial conditions
(position and velocity) of the bodies. Consider the motion of two bodies as shown in a). The masses of the
two bodies are M = 4 and m = 1; for convenience G was set equal to 10. The initial conditions (vector
components) are given as r m = (1, 0), r m = (2, 3) and r M = (2, 0) and r M = (2, 0). The motion of the
two bodies with time is shown in a). From the boundary conditions, we obtain the position of the center of
mass with time as r G = (7/5, 0) + (6/5, 3/5)t; this position with time is shown in b). The bodies orbit
about the instantaneous position of the center of mass.

The solution to the Kepler problem for these bodies is shown in c); the solution to the Kepler orbital
problem gives the instantaneous position of the relative position of the two bodies, r(t) = r m r m . The
Kepler problem has its origin as the center of mass, which also is the focus of the elliptical orbit. To recover
the orbits of the two bodies, we use equation (4). The two orbits are shown in d). These are also the
solutions that would be obtained by a direct numerical solution of the two-body problem with boundary
conditions chosen to place the center of mass at the origin. The origin serves as the focus for each elliptical
orbit. This example shows the importance of formulating the velocity and position boundary conditions so
that the center of mass remains xed at the origin. If this is done, the bodies will orbit about the center of
mass, producing the simplest solution to the two-body problem.

Equations of Motion
The equation of motion (F = ma), is

m
er = mr.
r2

Since the only force in the system is directed towards point O, the angular momentum of m with respect
to the origin will be constant. Therefore, the position and velocity vectors, r and r , will be in a plane

orthogonal to the angular momentum vector, and, as a consequence, the motion will be planar. Using
cylindrical coordinates, with ez being parallel to the angular momentum vector, we have,

er = (
r r2 )er + (r + 2r
)e .
r2

Now, we consider the radial and circumferential components of this equation separately.
Circumferential component
We have,
0 = r + 2r
.
Using the following identity,
1
r

d 2
(r ) = r + 2r,

dt

the above equation implies that


r2 = h constant.

(8)

We note that the constant of integration, h, that will be determined by the initial conditions, is precisely
the magnitude of the specic angular momentum vector, i.e. h = |r v|.
In a time dt, the area, dA, swept by r will be dA = r rd/2.

Therefore,
dA
1
h
= r2 = ,
dt
2
2
which proves Keplers second law:The line joining a planet to the Sun sweeps out equal areas in equal
intervals of time.
Radial component
The radial component of the equation of motion reads,

d
Since r2 dt

1
r

= r r2 .
r2

= r,
and r2 = h/ from equation 8, we can write


h d 1
d 1
r =
= h
.
d r
dt r

Dierentiating with respect to time,


r = h

d2
d2



1
h2 d2 1
= 2 2
.
r
r d
r
5

(9)

Inserting this expression into equation 9, and using equation 8, we obtain the following dierential equation
for 1/r as a function of .
d2
d2


1
1

+ = 2.
r
r
h

This is a linear second order ordinary dierential equation which has a general solution of the form,
1

= 2 (1 + e cos( + )) ,
r
h
where e and are two constants of integration. If we choose to be zero when r is minimum, then e will be
positive, and = 0. The equation describing the trajectory will be
r=

h2 /
.
1 + e cos

(10)

We shall see below that this is the equation of a conic section in polar coordinates.

Conic Sections
Conic sections are planar curves that are dened as follows: given a line, or directrix, and a point, or focus
O, a conic section is the locus of points, P , such that the ratio of the distance between the point and the
focus, P O, to the distance between the point and the directrix, P A, is a constant e. That is, e = P O/P A.

Since P O = r and P A = p/e r cos , we have


r=

p
.
1 + e cos

(11)

Here, p is the parameter of the conic and is equal to r when = 90o . The constant e 0 is called
the eccentricity, and, depending on its value, the conic surface will be either an open or closed curve. In
particular, we have that when
e=0

the curve is a circle

e<1

the curve is an ellipse

e=1

the curve is a parabola

e>1

the curve is a hyperbola.


6

Comparing equation(11) which deals solely with the property of a conic section, and equation(10) which
provides the solution of the motion of a point mass in a gravitational eld, we can identify the properties
of the conic section orbits in terms of the physical parameters of the Kepler problem. In particular, we see
that the trajectory of a mass under the inuence of a central force will be a conic curve with parameter
p = h2 /.

(12)

When e < 1, the trajectory is an ellipse, thus proving Keplers rst law:The orbits of the planets are
ellipses with the Sun at one focus.

The point in the trajectory which is closest to the focus is called the

periapsis and is denoted by . For elliptical orbits, the point in the trajectory which is farthest away from
the focus is called the apoapsis and is denoted by . When considering orbits around the earth, these points
are called the perigee and apogee, whereas for orbits around the sun, these points are called the perihelion
and aphelion, respectively.

Elliptical Trajectories
If a is the semi-major axis of the ellipse, then
2a = r + r .

(13)

Using equation 11 to evaluate r ( = 0) and r ( = ), we obtain


a = p/(1 e2 ).

(14)

Thus from the geometric properties of an ellipse,


r =

p
= a(1 e),
1+e

r =

p
= a(1 + e).
1e

Also, the distance between O and the center of the ellipse will be
a r = a e.

(15)

Other geometric properties of the ellipse are that the distance between point D and the directrix will be

equal to DO/e, which in turn will be equal to the sum of the distance between the focus and the center of

the ellipse, plus the distance between the focus and the directrix. That is, DO/e = ae + p/e. Therefore,

DO = a e2 + p = a. Hence, using Pythagoras theorem, b2 + (a e)2 = a2 , the semi-minor axis of the ellipse

will be b = a 1 e2 .

The area of the ellipse is given by

A = ab.

(16)

dA/dt = h/2

(17)

A = h /2,

(18)

Also, since

is a constant, we have

where is the period of the orbit. Equating these two expressions and expressing h in terms of the semi-major
axis as
h2 = p = a(1 e2 ),
we have

a3 ,

(19)

(20)

which proves Keplers third law:The square of the period of a planet is proportional to the cube of the
major axis of its elliptical orbit. This can be rewritten to obtain the time of ight or period of the orbit.
2
= a3/2

(21)

Time of Flight (TOF) in Elliptical Trajectories


We have found r(), the prediction of the shape of the orbit. However, this solution gives us no direct
information about the time behavior of the motions, such as (t). In many situations we will need to
determine the time of ight between two arbitrary points along the ellipse. In order to do that, we use
Keplers second law, which states that the motion of the planet sweeps out area at a constant rate.
Consider the orbital motion from point 0, to point 1. We would like to determine the time taken t1 . If the
motion continues, returning to point 2, the total time taken will be . We dene the time to reach point 1
as T1 and the time to reach point 2 as t2 . The total time taken is the t1 + t2 = , where is the total period
of the orbit.

From Keplers second law, equal areas are swept out in equal times. Thus the time taken to reach point 1
is given by
t1 /A1 = t2 /A2 = /ab

(22)

where ab is the total area of the ellipse, ab = A1 + A2 .

We now construct a more detailed analysis to determine the area Ap swept out by the orbit to a point P.

Referring to the gure, we see that the time required to travel between the point , the periapsis, and an

arbitrary point P is proportional to the curved area denoted by AP (AP is the sector dened by O, , P ).

More specically, since the total period of the orbit is and the total area of an ellipse is ab, tP , the time

required to travel from to P , equals the fraction that the area AP represents of the total area of the ellipse.

tP =

AP
.
ab

(23)

To nd the area AP we construct a circle of radius a with origin at the center of the ellipse. We identify a
point P on the circle to be in a vertical line with the point of interest P on the ellipse intersecting the point
O on the axis.

The various geometric quantities of the elliptical orbit have standard denitions: the position angle is often
called the true anomaly. The radial line of the circle for the origin O to P and the major axis of the ellipse
9

major axis dene an angle u, which is referred to as the eccentric anomaly. In addition, we dene a third
anomaly, the mean anomaly M of the point P , as
MP =

2tP
.

(24)

Here, tP is the time of ight from the periapsis to the point P .


Thus, if we want to determine the time of ight between two points 1 and 2 on the ellipse, we can use
equation (24) and write
T OF = t2 t1 =

A2 A1
(M2 M1 ) =
,
2
ab

where A2 A1 is the area swept out between points 1 and 2.


The mean anomaly for point P can also be written as
MP =

2 AP
,
AT

(25)

where AP is the area swept out up to the point P . When the area swept out equals the total area of the
ellipse AT , the time t equals the period and the mean anomaly M = 2 . (The subscript denotes
the return to the periapsis .) Thus the mean anomaly can be thought of as the fraction of the total angle
2 that would be swept out in a time by an object reaching point P . The focus is on time not on actual
spatial angle.
All is needed now is an expression for the mean anomaly M as a function of the orbit parameters. We start
by obtaining a relation between and u. From simple trigonometry, we have that
a cos u r cos = ae

(26)

or noting that r = a(1 e2 )/(1 + e cos ),


cos u =

e + cos
,
1 + e cos

cos =

cos u e
.
1 e cos u

(27)

We now develop relationships between the various areas indicated on the gure, with the goal to nd the
formula for the area AP , the area swept out by the point r as it travels from the periapsis to the point P .
The area A1 is the wedge in the circle occupied by the angle u. A1 = a2 u; the area of the large triangle
formed by the angle u within the circle is A2 = a2 cos(u)sin(u)/2. Therefore, the area of the large curved
segment from O, P , is
A1 A2 = (1/2) a2 (u Cos(u)Sin(u)).

(28)

The base of the small triangle of area A4 , O, O , P , is r cos() = a cos(u) ae by equation (26). The
height of the small triangle is b sin(u). Therefore, the area of the small triangle is A4 = (1/2) (a
cos(u) e) b sin(u). This area plus the curved segment O, P, is the total area swept by the point P .
The nal step in identifying the area segment swept out between point and P is to identify the curved
segment from O, P, , which is then added to the triangle section A4 to form the complete swept area.
The curved vertical segment formed from removing the large imbedded triangle A2 from the arc segment of
10

the circle A1 call it A3 is geometrically similar to the curved segment formed by removing the small
triangle from the area of the swept segment of the ellipse. Since the vertical height of the ellipse is b, and
the vertical height of the circle is a, the area of the desired curved segment can by obtained from that of the
corresponding segment of the circle by multiplying by b/a. Specically,
AP A4 = (b/a) (A1 A2 ).

(29)

Therefore, the nal result for the area swept out by the point r moving from point to point P is
AP = b/a (A1 A2 ) + A4

(30)

And the mean anomaly for the point P is


MP =

2 AP
2 (b/a (A1 A2 )) + A4
=
AT
AT

(31)

Thus, combining equations (24), (28) and (30), we obtain the mean anomaly for the point P , called Keplers
equation (It took a Kepler to work this out.)
u e sin u = MP =

2tP
.

where u is the eccentric anomaly for the point P , dened in the gure. This equation is very easy to use if
we want to know the time tP at which the satellite is at position . The only thing required, in this case,
is the calculation of the eccentric anomaly u using equation (27). On the other hand, if we need to nd
the position of the satellite at a given time t, then, we need to solve Keplers equation which is non-linear
using an iterative numerical algorithm such as Newtons method.

ADDITIONAL READING
J.L. Meriam and L.G. Kraige, Engineering Mechanics, DYNAMICS, 5th Edition
3/13 (except energy analysis)

11

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

J. Peraire, S. Widnall
16.07 Dynamics
Fall 2008
Version 1.1

Lecture L16 - Central Force Motion: Orbits


In lecture L12, we derived three basic relationships embodying Keplers laws:
Equation for the orbit trajectory,
h2 /
r=
1 + e cos

a(1 e2 )
=
.
1 + e cos

elliptical orbits

(1)

Conservation of angular momentum,


h = r2 = |r v| .

(2)

Relationship between the major semi-axis and the period of an elliptical orbit,

a3 .

(3)

Time of Flight (TOF) expressions for elliptical orbits,

(MB MA )
2
M = u e sin u
e + cos
cos u =
.
1 + e cos

T OF = tB tA

(4)
(5)
(6)

In this lecture, we will rst derive an additional useful relationship expressing conservation of energy, and
then examine dierent types of trajectories.

Energy Integral
Since there are no dissipative mechanisms and the only force acting on m can be derived from a gravitational
potential, the total energy for the orbit will be conserved. Recall that the gravitational potential per unit
mass is given by /r. That is, F /m = (/r) = (/r2 )er . Note that the origin (zero potential)
for the gravitational potential is taken to be at innity. Therefore, for nite values of r, the potential is
negative. The kinetic energy per unit mass is v 2 /2. Therefore,
1 2
v = E constant .
2
r

The total specic energy, E, can be related to the parameters dening the trajectory by evaluating the
total energy at the orbits periapsis ( = 0). From equation 1, r = (h2 /)/(1 + e), and, from equation 2,
v2 = h2 /r2 = (1 + e)/r , since r and v are orthogonal at the periapsis. Thus,
E=

1 2

2
v
= 2 (e2 1).
2
2h
r

(7)

We see that the value of the eccentricity determines the sign of E. In particular, for
e<1

the trajectory is closed (ellipse), and

E < 0,

e=1

the trajectory is open (parabola), and

E = 0,

e>1

the trajectory is open (hyperbola), and

E > 0.

Equations 1, 2, and 3, together with the energy integral 7, provide most of relationships necessary to solve
basic engineering problems in orbital mechanics.

Types of Orbits
Elliptic Orbits (e < 1)
When the trajectory is elliptical, h2 = a(1 e2 ) (see lecture L12). Then, the total specic energy simplies
to E = /(2a), and the conservation of energy can be expressed as
1 2

v = .
2
r
2a

(8)

This expression shows that the energy (and the period) of an elliptical orbit depends only on the major
semi-axis. We also see that for a xed a, the value of h determines the eccentricity. There are two limiting
cases: e 1, which gives h 0, which in turn implies that the minor semi-axis of the ellipse b 0; and

e = 0 which corresponds to a circular orbit with h = a. In the rst case, the maximum value of the
eccentricity is limited by the size of the planet, since, for suciently large values of e, the trajectory will
collapse onto the planets surface.
Below we show three elliptical trajectories that have the same energy (same value of a), but dierent
eccentricities.

Circular Orbits (e = 0)
This is a particular case of an elliptic orbit. The energy equation is given by equation 8. The radius is
constant
h2
v2 r2
= c .

For orbits around the earth, = gR2 , where g is the acceleration of gravity at the earths surface, and R is
r=

the radius of the earth. Then,

gR2
=
,
(9)
r
r
which shows that the velocity of a circular orbit is inversely proportional to the radius. We now consider
vc2 =

two particular orbits of interest:


1) r = R
This corresponds to a hypothetical satellite orbiting the earth at a zero altitude above the earths
surface. The orbits velocity is
vc =

gR = 7910 m/s,

and the period, from equation 3, is


2
=
R3/2 = 2
gR2

R
= 84.4 min .
g

This period is called the Schuler period, and it is the minimum period that any free ight object can
have in orbit around the earth.
2) Synchronous Orbits
These are orbits whose period is the same as the earths rotational period ( 24 h). In addition, if the
orbit is in the equatorial plane, the orbit is said to be geostationary because the satellite will stay xed
relative to an observer on the earth. Using equation 3,
2 2 1/3
gR
a=
= 42042 km 6.6R,
4 2

which corresponds to an altitude above the surface of 5.6R.

Example

Elliptical Orbits

Consider a satellite launched from an altitude d above the earths surface, with velocity vc =

/(R + d).

If the direction of the velocity is orthogonal to the position vector, the trajectory will clearly be a circular
orbit of radius R + d. However, if the velocity is in any other direction, the trajectory will be an ellipse of
semi-major axis equal to R+d. The characteristics of the ellipse can easily be determined as follows: knowing
r and v, we can determine h; using equation 7, we can determine e; and from the trajectory equation 1, we
can determine , and hence the orientation of the ellipse.

Parabolic Orbits (e = 1)
From equation 1, we see that r for . From the energy integral, with E = 0, we have that,
1 2
v = 0,
2 e
r

ve2 =

2
.
r

(10)

Here, ve is the escape velocity the smallest velocity needed to escape the eld of gravitational attraction.

Comparing equations 9 and 10, we see that, for a given r, the escape velocity is a factor of 2 larger than

the velocity necessary to maintain a circular orbit. Thus, if a satellite is on a circular orbit with velocity vc ,

the necessary v to escape is ( 2 1)vc .

It should be noted that a satellite in a parabolic trajectory has a total specic energy, E, equal to zero. This

means that when r increases, the kinetic energy is transformed to potential energy such that, at innity, the

residual velocity is equal to zero.

Hyperbolic Trajectory (e > 1)


For a hyperbolic orbit, e > 1 and the semimajor axis a is negative. The energy is constant and given by

E=

v2

v2
=
=
2a
2
r
2

(11)

Therefore the magnitude of the velocity inbound is the same as the velocity outbound.
Hyperbolic orbits are the linkages between orbits about a given planet and interplanetary travel. The
geometry of a given hyperbolic orbit is shown below (Figure below taken from Kaplan.) This basic orbit is
used to describe a planetary yby and/or a hyperbolic escape from a planetary orbit.
For a given case, the point rp is take as a point on a planetary orbit, and the velocity vp is taken as the
velocity of a satellite required to escape on a given hyperbolic trajectory from a circular orbit passing through
the point rp . From this, the trajectory is determined including v and .

Geometry of hyperbolic passage.

+
v

Asymptote
( - )
2
2

rp

P
-a

Image by MIT OpenCourseWare.

This solution can be used to represent a yby of a spacecraft past a planet. The direction of the orbit
changes due to the yby but the velocity v (and energy) does not. However, this solution to the 2-body
orbit problem is with respect to coordinates xed at the center of mass; for a satellite-planet orbit, this is
a coordinate system moving with the planet. Although the inbound and outbound velocities are the same
with respect to the moving planet, when this is viewed in an inertial reference frame, considerable
change in the satellite velocity can occur as a result of a yby. This eect has been widely used in the design
of planetary missions.
From the general orbit equation, valid for all values of ,
r=

a(1 2 )
1 + cos

(12)

when cos1 (1/e), we have r . Hence, the trajectories are open. Moreover, if the velocity
v, at a given r is known (such as near the planet), from conservation of energy the velocity at innity is

simply v = 2E = 2v 2 2/r. For a given energy level, the eccentricity of the orbit is determined by

h = v =

2 (2 1)
.
2
v

h is a constant of the motion. This yields


2 = 1 +

4
v
2
2

(13)

Another useful form is obtained by expressing variables at the point of closest approach, the periapsis.
=1+

2
rp v

(14)

From conservation of angular momentum, we obtain the displacement of the trajectory from a parallel
line through the center of the planet.

(2 1)
=
2
v

(15)

An important parameter for hyperbolic orbits is the turning angle, , which is the angle through which the
velocity changes along the trajectory as the body travels from to . The turning angle is given by
= 2( ), or = 2 sin1 (1/e). Below we show several hyperbolic trajectories which have identical
terminal velocities for dierent values of the eccentricity (and turning angle).

Example

Dierent orbits as a function of v0

We consider the problem of launching a satellite at an altitude d with an initial velocity v0 , along the
direction tangent to the earths surface. We consider the dierent trajectories that are obtained as we vary
the magnitude of v0 .

For v0 = v0c
/(R + d), the trajectory will be a circle (e = 0). For v0 = v0e
2/(R + d),
the trajectory will be parabola (e = 1). For v0 > v0e , the trajectory will be a hyperbola, whereas for
v0c < v0 < v0e the trajectory will be elliptical. We note that, for all these orbits, the launch point, P , is the
orbits perigee, or the closest point in the trajectory to the earths center.

On the other hand, when the velocity v0 < v0c , the straightforward use of expressions 1 and 2 gives a negative
eccentricity! The eccentricity is negative because equation 1 assumes that the origin of is taken to be at
the orbits perigee. In turns out that in this case, the orbit has a lower energy than the circular orbit, and,
hence, the launch point is now the orbits apogee. The proper use of equation 1 requires that = . In this
case, we have d + R = (v02 (d + R)2 /)/(1 + e cos ), which for v0 < v0c gives a positive eccentricity. In the
picture above, we see one such trajectory depicted with a dotted line.

ADDITIONAL READING
J.L. Meriam and L.G. Kraige, Engineering Mechanics, DYNAMICS, 5th Edition
3/13 (energy analysis)

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

S. Widnall, J. Peraire
16.07 Dynamics
Fall 2008
Version 2.0

Lecture L17 - Orbit Transfers and Interplanetary Trajectories


In this lecture, we will consider how to transfer from one orbit, to another or to construct an interplanetary
trajectory. One of the assumptions that we shall make is that the velocity changes of the spacecraft, due
to the propulsive eects, occur instantaneously. Although it obviously takes some time for the spacecraft to
accelerate to the velocity of the new orbit, this assumption is reasonable when the burn time of the rocket
is much smaller than the period of the orbit. In such cases, the v required to do the maneuver is simply
the dierence between the velocity of the nal orbit minus the velocity of the initial orbit.
When the initial and nal orbits intersect, the transfer can be accomplished with a single impulse. For more
general cases, multiple impulses and intermediate transfer orbits may be required.
Given initial and nal orbits, the objective is generally to perform the transfer with a minimum v. In some
situations, however, the time needed to complete the transfer may also be an important consideration.
Most orbit transfers will require a change in the orbits total specic energy, E. Let us consider the change
in total energy obtained by an instantaneous impulse v. If v i is the initial velocity, the nal velocity, v f ,
will simply be,
v f = v i + v .

If we now look at the magnitude of these vectors, we have,


vf2 = vi2 + v 2 + 2vi v cos ,
where is the angle between v i and v. The energy change will be
E =

1 2
v + vi v cos .
2

From this expression, we conclude that, for a given v, the change in energy will be largest when:
- v i and v are co-linear ( = 0), and,
- vi is maximum.

For example, to transfer a satellite on an elliptical orbit to an escape trajectory, the most energy ecient
impulse would be co-linear with the velocity and applied at the instant when the satellite is at the elliptical
orbits perigee, since at that point, the velocity is maximum. Of course, for many required maneuvers, the
applied impulses are such that they cannot satisfy one or both of the above conditions. For instance, ring
at the perigee in the previous example may cause the satellite to escape in a particular direction which may
not be the required one.

Hohmann Transfer
A Hohmann Transfer is a two-impulse elliptical transfer between two co-planar circular orbits. The transfer
itself consists of an elliptical orbit with a perigee at the inner orbit and an apogee at the outer orbit.
The fundamental assumption behind the Hohmann transfer, is that there is only one body which exerts
a gravitational force on the body of interest, such as a satellite. This is a good model for transferring an
earth-based satellite from a low orbit to say a geosynchronous orbit. Inherent in the model is that there is
no additional body sharing the orbit which could induce a gravitational attraction on the body of interest.
Thus, as we shall discuss, the Hohmann transfer is a good model for the outer trajectory of an earth-Mars
transfer, but we must pay some attention to escaping the earths gravitational eld before were on our
way.

It turns out that this transfer is usually optimal, as it requires the minimum vT = |v |+|v | to perform
a transfer between two circular orbits. The exception for which Hohmann transfers are not optimal is for
very large ratios of r2 /r1 , as discussed below.
The transfer orbit has a semi-major axis, a, which is
a=

r1 + r2
.
2
2

Hence, the energy of the transfer orbit is greater than the energy of the inner orbit (a = r1 ), and smaller
than the energy of the outer orbit (a = r2 ). The velocities of the transfer orbit at perigee and apogee are
given, from the conservation of energy equation, as

2
2
2
v =

(1)
r1
r1 + r2

2
2
v2 =

.
(2)
r2
r1 + r2

The velocities of the circular orbits are vc1 = /r1 and vc2 = /r2 . Hence, the required impulses at
perigee and apogee are,

2r2
= v vc1 =
1
r1 + r2

2r1
= vc2 v =
1
.
r1 + r2
r2

v
v

r1

If the initial orbit has a radius larger than the nal orbit, the same strategy can be followed but in this case,
negative impulses will be required, rst at apogee and then at perigee, to decelerate the satellite.

Example

Hohmann transfer [1]

A communication satellite was carried by the Space Shuttle into low earth orbit (LEO) at an altitude of 322
km and is to be transferred to a geostationary orbit (GEO) at 35, 860 km using a Hohmann transfer. The
characteristics of the transfer ellipse and the total v required, vT , can be determined as follows:
For the inner orbit, we have,
r1
vc1
E1

= R + d1 = 6.378 106 + 322 103 = 6.70 106 m

gR2
=
=
= 7713 m/s
r1
r1
=

gR2
=
= 2.975 107 m2 /s2 (J/kg).
2r1
2r1

Similarly, for the outer circular orbit,


r2

= R + d2 = 42.24 106 m

vc2

= 3072 m/s

E2

= 4.718 106 m2 /s2 (J/kg).

For the transfer trajectory,


2a = r1 + r2 = 48.94 106 m

E =
= 8.144 106 m2 /s2 (J/kg),
2a

which shows that E1 < E < E2 . The velocity of the elliptical transfer orbit at the perigee and apogee can
be determined from equations 3 and 4, as,
v

= 10, 130 m/s ,

1, 067 m/s .

Since the velocity at the perigee is orthogonal to the position vector, the specic angular momentum of the
transfer orbit is,
h = r1 v = 6.787 1010 m2 /s ,
and the eccentricity can be determined as,

e=

1+

2Eh2
= 0.7265 .
2

Finally, the impulses required are,


v

= v vc1 = 10, 130 7713 = 2414 m/s

= vc2 v = 3072 1607 = 1465 m/s,

The sign of the vs indicates the direction of thrusting (whether the energy is to be increased or decreased)
and the total v is the sum of the magnitudes. Thus,
vT = |v | + |v | = 2417 + 1465 = 3882 m/s.
Since the transfer trajectory is one half of an ellipse, the time of ight (TOF) is simply half of the period,

(24.47 106 )3
T OF =
= 19, 050 s = 5.29 h .
3.986 1014
In order to illustrate the optimal nature of the Hohmann transfer, we consider now an alternative transfer
in which we arbitrarily double the value of the semi-major axis of the Hohmann transfer ellipse, and nd
the characteristics and vT of the resulting fast transfer. The semi-major axis of the transfer ellipse will
be 2a = 98 106 m, and E = /(2a) = 4.067 106 m2 /s2 (J/kg). The velocity of the transfer orbit at
departure will be

3.986 1014
6
v = 2 4.067 10 +
= 10.530 m/s,
6.70 106
and,
v = 10, 530 7713 = 2817 m/s .
The specic angular momentum is,
h = 10, 530 (6.70 106 ) = 7.055 1010 m2 /s,
and the orbits eccentricity, e, is 0.863.
4

We can now calculate, from the energy conservation equation, the velocity of the transfer orbit at the point
of interception with the outer orbit, vint ,

vint =

2 4.067 106 +

3.986 1014
42.24 106

= 3277 m/s .

Since the angular momentum, h, is conserved, we can determine the component of v int in the circumferential
direction
(v int ) =

h
= 1670 m/s
r2

and the elevation angle, , is thus,


= cos1

(v int )
= 59.36o
vint

Finally, from geometrical considerations,


2
2
2
vint
= vc2
+ vint
2vc2 vint cos ,

which yields vint = 3142 m/s, so that


vT = 2817 + 3142 = 5959 m/s.
Comparing to the value of the Hohmann transfer vT of 3875 m/s, we see that the fast transfer requires
a vT which is 54% higher. The analysis of elliptical trajectories which intersect the circular orbits at an
angle is referred to as Lamberts problem. These trajectories can have faster transit times but at a greater
cost in energy.
It can be shown that when the separation between the inner and outer orbits is very large (r2 > 11.9r1 )
(a situation which rarely occurs), a three impulse transfer comprising of two ellipses can be more energy
ecient than a two-impulse Hohmann transfer. This transfer is illustrated in the picture below. Notice that
the distance from the origin at which the two transfer ellipses intersect is a free parameter, which can be
determined to minimize the total v. Notice also that the nal impulse is a v which opposes direction of
motion, in order to decelerate from the large energy ellipse to the nal circular orbit. Although this transfer
5

may be more energy ecient relative to the two-impulse Hohmann transfer, it often involves much larger
travel times. Try it!

Interplanetary Transfers
The ideas of Hohmann transfer can be applied to interplanetary transfers with some modication. The
Hohmann transfer for satellite orbits assumes the satellite is in a circular orbit about a central body and
desires to transfer to another circular and coplanar orbit about the central body. It also assumes that no
other gravitational inuence is nearby. When more than one planet is involved, such as a satellite in earth
orbit which desires to transfer via a Hohmann orbit about the sun to an orbit about another planet, such as
done in a mission to Mars. In this case, the problem is no longer a two-body problem.
Nevertheless, it is common (at least to get a good approximation) to decompose the problem into a series of
two body problems. Consider, for example, an interplanetary transfer in the solar system. For each planet
we dene the sphere of inuence (SOI). Essentially, this is the region where the gravitational attraction
due to the planet is larger than that of the sun. In order to be on our way to the destination planet, we
must climb out of the potential well of the originating planet. We will use a hyperbolic escape orbit to
accomplish this. Alternatively, we could do a direct calculation including the position of all of the bodies:
the sun, earth and Mars. However, because the time scale and length scales are so dierent for the dierent
phases of the mission, it requires special attention to the details of the numerical method to attain good
accuracy. The method of patched conics is a good place to start our analysis.

The mission is broken into phases that are connected by patches where each patch is the solution of a two
body problem. This is called the patched conic approach. Consider, for instance, a mission to Mars. The
rst phase will consist of a geocentric hyperbola as the spacecraft escapes from earth SOI, attaining a velocity
v1 in a direction beyond the earths SOI. The second phase would start at the edge of the earths SOI,
and would be an elliptical trajectory around the sun while the spacecraft travels to Mars. This orbit could
be part of a Hohmann transfer sequence; in this case, v1 would be the Hohmann transfer velocity after the
V has been applied. The third phase would start at the edge of Mars SOI, and would be a hyperbolic
approach capture trajectory with the gravitational eld of Mars as the attracting force. This third phase
can be thought of as a combination of a Hohmann transfer and a hyperbolic capture by the planet.
The time scales and length scales for the various phases of the mission are quite dierent. The time for a
transfer to another planet is measured in months or years; the time scale for escape for a planet is measured
in days or hours. The length scale for planetary trajectories is measured in AU units where AU is the distance
from the earth to the sun; the length scale for hyperbolic escape from a planet is measure in distances typical
of planetary radii and orbits. This can be a challenge for an orbit calculation program since the step size
must change dramatically near a planet. We will examine this problem analytically using the method of
matched conics to get an approximate result. This approach does not work well for trajectories from earth
to moon since the moon is in the SOI of the earth.
In turns out that a Hohmann transfer orbit with hyperbolic escape and capture trajectories can be shown to
be the minimum energy trajectory for a planetary mission just as the Hohmann transfer itself is an optimum
solution for a specied satellite change of orbit. Of course a successful mission requires that the time of
launch be selected so that the desired destination planet is at the destination when the spacecraft arrives.
We assume that the proper launch time has been selected.

Hyperbolic Escape
We now consider a hyperbolic escape from a planetary orbit of a planet in orbit about the sun into an
elliptical orbit about the sun determined by a Hohmann transfer to the desired planet combined with a
hyperbolic capture into a planetary orbit about the destination planet in orbit about the sun for a complete
interplanetary transfer. This sequence is appropriate for travel to an outer planet such as Mars.
We rst consider the Hohmann transfer from a circular orbit about the sun coincident with the original
planetary orbit to a circular coplanar orbit coincident with the destination planets orbit about the sun. We
neglect the gravitational elds of the two planets in this calculations. From Equations (1-4) we have v and
v for the elliptical connecting orbit and
v2
v2

2
2
=

r1
r 1 + r2

2
2
=

r2
r1 + r2

(3)
(4)

and for the circular planetary orbits at r1 and r2 (Note that we have used the symbol without designation.
Obviously, the choice of in a particular calculation depends upon the celestial body supporting the orbit.)

v1 = /r1 and v2 = /r2 for the initial and nal circular orbits about the sun.

This determines the v and v for the Hohmann interplanetary transfer. It also determines the initial
and nal conditions for a hyperbolic escape trajectory.
Consider conditions at the departure planet, denoted by subscript 1. The planet moves with the circular
orbit velocity v1 . The hyperbolic trajectory is dened in a coordinate system relative to the moving planet,
and the escape trajectory must end with the velocity appropriate for the Hohmann transfer ellipse in inertial
space, v or v v1 relative to the planet. The hyperbolic escape velocity v then equals exactly the v
for a traditional Hohmann transfer. The only remaining unknown is the condition of the spacecraft when it
decides to escape.
8

The geometry of a hyperbolic escape comes directly from the general geometry of a hyperbolic orbit. In this
case we desired to depart on a vertical trajectory ending with a velocity v = v v1 . If the spacecraft is
in orbit around the earth at a velocity vorbit and a radius r1 , this requires a vescape = v1p vorbit where
v1p is the velocity of the hyperbolic escape orbit at the periapsis rp = r1 .
Since energy is constant,

v12p

v2

= 1 .
2
2
r1p

(5)

The parameters of the hyperbolic escape orbit follow from these conditions. The eccentricity is given by
1 = 1 +

2
r1p v1

(6)

and the angular change from periapsis to innity, 1 /2, is given by


1 /2 = sin1

1
1

(7)

For the orbit geometry sketched below, this means that the launch from a circular orbit into a hyperbolic
orbit with a nal vertical direction as sketched occurs at an angle 1 /2 as shown in the gure.

The trajectory shown is a counterclockwise escape trajectory. A clockwise escape trajectory is also possible
but, because of the counterclockwise rotation of the earth, counterclockwise is preferred. One might worry
about the sidewise displacement 1 of the trajectory from the centerline of the planet orbit. However,
this distance is very small compared to the interplanetary distances dened by the distance to the sun and
is typical of the small errors inherent in the method of patched conics. Therefore, the hyperbolic escape
trajectory has been dened as shown below. This enables the determination of the v1 for this portion of
the mission. The hyperbolic escape trajectory is shown below.

Hyperbolic Capture
Also shown below is the hyperbolic capture trajectory at the nal planetary destination. Both these trajec
tories are shown for the case of travel to an outer planet such as Mars. For travel to an inner planet such as
Venus, the roles of escape and capture would be reversed.

We now consider the hyperbolic capture trajectory. From the Hohmann transfer calculation, we obtain the
result that an additional v in the ight direction is required to circularize the orbit of the space craft
into the destination planetary orbit, subscript 2. This means that the velocity v of the elliptical Hohmann
transfer orbit is less that the velocity of the planet in its circular orbit. Thus the planet will overtake the
spacecraft resulting in a hyperbolic capture orbit. The velocity v2 as seen by the planet is v v2 and is
directed towards the planet; this quantity plays the role of v2 in the hyperbolic capture orbit. We desire to
capture the planet into an orbit of radius r2p and inquire what conditions are require to achieve this. This
will determine 2 and v2p . From conservation of energy, we have v22p /2 /r2p = v22 ; 2 = 1 + r2p v22 /;
and sin2 /2 = 1/2 . Of course some mid-course correction may be required to set 2 to achieve this capture.
This will be ignored.
The nal step in the interplanetary mission would be to insert the spacecraft into a circular orbit of radius
r2p . We have chosen to capture the spacecraft in a planetary orbit proceeding in the direction of the planets
rotation about its axis. If we only intend to stay in orbit, this doesnt matter. However, if we intend to
descend and later return to orbit, these maneuvers should be designed with the planets rotation in mind,
jus as orbits from earth take advantage of the direction of the earths rotation to save energy. We can easily
determine the v required to convert the hyperbolic capture orbit into an orbit about the capturing planet.
10

This mission analysis is a bit arbitrary; the spacecraft could also descend to the planet surface or yby to
another destination. But if we desire to calculate the energy requirements for an earth to Mars mission, this
is a reasonable formulation and allows a comparison between several options including the more complex
trajectories using solutions to Lamberts problem. We can also complete a round-trip mission by remaining
in orbit for a specied time and then retracing the process to return to earth. The v s required to descend
to and depart from the planet surface could also be included. These combined Hohmann, escape and capture
trajectories are minimum energy trajectories for a given mission and serve as useful benchmarks. When the
orbits of the two planets are no longer co-planar, even by a small angle such as the 1.8o of Mars/earth
orbits, the Hohmann transfer is no longer optimum, or even practical, since the plane containing the sun
and the two planets at departure and arrival for a Hohmann transfer is substantially tipped with respect to
the planetary orbits.
For missions designed to visit several planets, the situation can become very complex as one often tries to
take advantage of the gravitational elds of the planets encountered on the way, by entering into their SOIs
with the objective of either changing direction or gaining additional impulse. This technique is often referred
to as gravity assist or yby.

References
[1] F.J. Hale, Introduction to Space Flight, Prentice-Hall, 1994.
[2] M.H. Kaplan, Modern Spacecraft Dynamics and Control, John Wiley and Sons, 1976.

11

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

S. Widnall
16.07 Dynamics
Fall 2008
Version 1.0

Lecture L18 - Exploring the Neighborhood: the Restricted

Three-Body Problem

The Three-Body Problem


In Lecture 15-17, we presented the solution to the two-body problem for mutual gravitational attraction.
We were able to obtain this solution in closed form. For bodies of comparable mass, the solution showed
that the bodies orbited about their center of mass, with orbital periods that could be obtained analytically
by transforming the two-body problem into a Kepler problem with modied gravitational forces. For a
body of negligible mass, such as a satellite orbiting the earth, we were able to place the earth at rest at the
center of our coordinates, and obtain a simple expression for the orbit of the satellite.
No such solution is available for the three-body problem. The general solution to the three-body problem
must be obtained numerically, and even with these tools, the solution is chaotic: that is small changes in
initial conditional lead to widely diverging behavior.
For analysis of space missions, such a spacecraft moving between planets, or from earth to the moon, we
may apply the model known as the restricted three-body problem, in which the mass of a small spacecraft
is aected by the gravitational forces from two bodies, but the large bodies do not feel the inuence of the
spacecraft.
The restricted three-body problem is most easily analyized/understood in a coordinate system rotating with
the two primary bodies. Since the solution of the two-body problem is that of the two primary masses
rotating at constant angular velocity about their center of mass, in a coordinate system rotating with
angular velocity , these two bodies are stationary. The third body moves in their gravitational elds and
experiences the additional acceleration terms associated with the rotation of the coordinate system.
Consider a two-body system of mass M1 and M2 in an inertial references frame. We add a small mass m3
which does not aect the large masses.

We consider planar x,y motion only. The equations of motion are

x
1

x
2

x
3

y1

y2

y3

GM2 (x1 x2 )

(1)

(3/2)

((x1 x2 )2 + (y1 y2 )2 )
GM1 (x2 x1 )

(2)

(3/2)

((x2 x1 )2 + (y2 y1 )2 )
GM1 (x3 x1 )
)2

((x3 x1 + (y3 y1
GM2 (y1 y2 )

(3/2)
)2 )

GM2 (x3 x2 )
(

((x3 x2 )2 + (y3 y2 )2 ) 3/2)

(4)

(3/2)

((x1 x2 )2 + (y1 y2 )2 )
GM1 (y2 y1 )

(5)

(3/2)

((x2 x1 )2 + (y2 y1 )2 )
GM1 (y3 y1 )
((x3 x1 )2 + (y3 y1

(3/2)
)2 )

(3)

GM2 (y3 y2 )
(

((x3 x2 )2 + (y3 y2 )2 ) 3/2)

(6)

The solution to the motion of M1 and M2 is a Kepler problem, where the two bodies move in circular orbits
about their center of mass with a frequency

G(M1 + M2 )
,
R3

(7)

where R is the distance between them. R = r1 + r2 (r1 is negative). We can use these equations to solve
numerically for the motion of the small mass m3 in the presence of the gravitational eld of M1 and M2 ; in
general no analytic solution is possible.
However, considerable insight can be gained about the behavior of the small mass m3 by moving into a
coordinate system in which the two large masses M1 and M2 are stationary. Since the solution to the Kepler
problem for these two masses is a rotation of angular velocity about the center of mass, if we place the
origin of our coordinate system at the center of mass and rotate it with angular velocity , we will see that
M1 and M2 are at rest; their velocities v1 and v2 which are obtained from the solution to the Kepler problem
are exactly canceled by the eect of rotation, in which the velocity at r1 and r2 are canceled since r2 and
r1 are subtracted from v2 and v1 as the eect of the rotating coordinate system.
2

However, the acceleration in inertial space x


and y must be modied to account for the rotating coordinate
system. From Lecture 8, equations(9), and (10) the acceleration for planar motion x (t) and y (t) in a
rotating coordinate system is given by
x
(t)

y(t)

z(t)

d2 x
dy
x 2 2
2
dt
dt

d2 y
dx
2

+
2
dt2
dt
2
d z
dt2

(8)
(9)
(10)

Here we have included the acceleration in the z (out of plane) direction since it is unmodied by the rotation
of the coordinate system. This will allow us to study out of plane motions if we wish.
We now modify the equations for the motion of the three masses in their mutual gravitational eld by
modifying the acceleration according to the formulae given in equation (8), (9), and (10). The gravitational
terms are unmodied since they are unaected by the rotation of the coordinate system. (The gravitational
force have no time behavior and are unaected by relative motion.) However, only the motion of the small
mass m3 is of interest, since by the construction of the problem, M1 and M2 are stationary in our coordinate
system. We include the z(t) equation which is very straightforward; z(t) is unaected by the rotation of the
coordinate system; by our problem statement, the motion of the small mass does not aect the motion of
the two large masses; and the motion of the two large masses is planar, so that z1 and z2 are always zero.
d2 x3
dy
x3 2 2 3
2
dt
dt
d2 y3
dx
y3 2 + 2 3
2
dt
dt
d2 z3
dt2

GM1 (x3 x1 )

GM2 (x3 x2 )

(11)
(
((x3 x2 )2 + (y3 y2 )2 + z32 ) 3/2)
GM2 (y3 y2 )
=

(3/2)
3/2
((x3 x1 )2 + (y3 y1 )2 + z32 )
((x3 x2 )2 + (y3 y2 )2 + z32 )
GM1 (z3 )
GM2 (z3 )
=

3/2
3/2
((x3 x2 )2 + (y3 y2 )2 + (z3 )2 )
((x3 x2 )2 + (y3 y2 )2 + (z3 )2 )
3/2

((x3 x1 )2 + (y3 y1 )2 + z32 )


GM1 (y3 y1 )

We now drop the s for convenience, concentrate on the motion of m3 in the x, y plane, and rewrite the
equations moving the centripetal acceleration to the right side. The centripetal acceleration x2i
3

y2j can be written as the gradient of a scaler: x2i 2j = ( 12 2 r2 ). If we do this, we can see
that the entire right side of the equation can be written as the gradient of a scaler function of r, and thus
could be thought of as a potential, related to a conservative force eld as Fc = V (r). It contains both
the gravitational potential and the potential formed from the centrifugal force. Because we have left the
Coriolis acceleration on the left side of the equation, we do not have a simple equation for the behavior of a
particle in a conservative force eld in inertial space. However, if we are interested in the spatial behavior
of the force eld seen by a particle at rest, this formulation gives great insights. We write the equation for
the motion of m3 in the x, y plane as
dr ) = (Vg (r) 1 2 r2 )
r + 2(
dt
2

(12)

where Vg (r) is the gravitational potential for the two bodies: M1 and M2 located at x = r1 , y = 0 and
x = r2 , y = 0, as shown in the gures.
From Lecture 13, we have that the gravitational potenial/per unit mass of m3 is
GM1
GM2
Vg (r) =

(x + r1)2 + y 2
(x r2)2 + y 2

(13)

so that the total potential is


GM1
GM2
1
V (r) =

2 r 2
2
2
2
2
2
(x + r1) + y
(x r2) + y

(14)

where the eect of centrifugal force has been written as a potential, V = 12 2 r2 . We shall refer to the
combination of gravitational and centrifugal potential as a pseudo-potential, since when the velocity is not
zero, the additional Coriolis term must be included to determine the dynamics of the particle; in this case
we do not have the simple picture of a particle exchanging energy between potential and kinetic. We will
use it to investigate the neighborhood, and specically to ask whether there are any points where a particle
might be placed and experience no force.
For convenience, we set G = 1, M1 = 10, M2 = 1, and R = 10. This gives r1 = 10/11 and r2 = 100/11

and = 11/1000. We rst examine the gravitational potential given by equation (15) along the x axis
in comparison with the pseudo-potential. The result are shown in the gure. We recall, that at any point,
F = V . Although we are only looking at the behavior along y = 0, we see an interesting dierences
between the behavior of the gravitational potential and the pseudo-potential.

For the gravity potential, we do see a point where the gravitational pull of the two bodies is equal and
opposite (Think of the earth-moon system.). However, for the pseudo-potential, we see three points where
the gravitational plus centrifugal forces balance. To explore this further, we need to move into two dimensions
in the plane. The result for the gravitational potential alone is not meaningful, since we are in a rotating
coordinate system. The centrifugal force is a real force and must be included in any discussion of force;
however the gravitational potential does give some physical insight into whats going on. We now move
into two dimensions and examine the contour plot of the gravitational and the centrifugal potential; lines of
constant potential in the x, y plane.

The gure shows lines of constant potential for the gravitational potential of the two bodies, M1 and M2 .

We also show the potential for the centrifugal term which is a simple - 12 2 r2 term; lines of constant potential

are circles.
Now we combine these expressions to form the total potential for the two-body problem in a rotating
coordinate system in which the two bodies are at rest. This will allow us to answer the question: for a small
body m3 wandering about this neighborhood, are there any points where no force will be exerted on m3 and
therefore points at which m3 might be able to remain at rest under the action of a balance of forces. There
are ve such points called the Lagrange points. They play an important role in the theory of planetary
motion, and the design of space missions.

We see several regions of this gure which bear further examination. Recall from our examination of the
behavior of the pseudo-potential along the x axis, we identify three possible points where no force in the x
direction would be exerted on a particle. In fact, due to the symmetry of the potential in the y direction,
no y force will be felt on a particle located on the x axis. Therefore these three points are equilibrium
points, in the sense that no force acts upon a particle at these points. Recall that following any discussion
of equilibrium, is a discussion of stability: a ball balanced upon an overturned bowl is in equilibrium but
not stable; a ball at the bottom and inside a bowl is at equilibrium and stable. These three points are
visible in the contour plot, although the point outboard of M1 is indistinct; it is there. Before completing
our discussion, we turn briey to the closed contours o of the x axis. If this contour contains a bowl or
a potential well, we have a possible point of equilibrium. For a closed contour of a continuous function, we
are guaranteed a point of zero slope somewhere within the contour, a possible point of equilibrium; however,
our dynamics includes Coriolis forces which could aect the stability. We now analyze the location of these
points in more detail and greater generality.

We now return to the two-body problem and do not choose the particular numerical values for M1 , M2 and
R. We do take G = 1 as before (you can think of this a particular non-dimensionalization of the equations;
in any case from the form of Vg and the form of , G drops out. We introduce
r
r1

where =

M2
M1 +M2 ;

M1
M1 +M2

and =

r2

= x(t)i + y(t)j

(15)

= R

(16)

(17)

G(M1 +M2 )
R3

as required by the solution of the two-body problem.

With this formulation, we seek solutions for the points x and y where the force on a particle would be zero.
This is a complex problem, involving nding the roots of three fth-order polynomials. We could do such
a solution numerically, but for greater insight we seek an analytic solution. Such a solution can easily be
found for small , where is the mass ratio. In fact the existence of such points requires that be small,
so this is not a signicant restriction on the utility of the result. Expanding the total x and y forces on the
particle for small and setting the result equal to zero gives the following approximations for the location
of these three equilibrium points along the x axis. They are called Lagrange points L1 , L2 and L3 .

1/3
L1 = R 1
3

1/3
L2 = R 1 +
3


5
L3 = R 1 +
12
Identifying the remaining Lagrange point requires that we nd a force balance for y = 0.
7

(18)
(19)
(20)

Referring to the gure, we note that since the centrifugal force is aligned with vector r, a force balanceif
indeed there is onewould involve only gravitational terms. Thus in a direction perpendicular to r

Fper = R
resulting in a solution x =

1
2 (R

(x R)2 + y 2 )3/2
(x + R)2 + y 2 )3/2

=0

(21)

R): that is the x location of the Lagrange pointif there is one, lies

exactly half way between the two masses. With the result, the equation for the forces parallel to r reduce to

(x2 + y 2 )
1
1
Fpar = 2

(22)
R
R3
(x R)2 + y 2 )3/2
Since this result equates the distance between the point in question to the magnitude of R, we conclude
that the Lagrange point is at a distance R from M2 and by symmetry from M1 : L4 is at positive y; L5 its
mirror image. Since the masses are R apart on the x axis, the location of L4 is at the peak of an equilateral
triangle: R on all three sides. We add these points to the gure showing location of the Lagrange points.
This is a remarkable result. The location of the L4 and L5 , which have both x and y positions are


R (M1 M2 )
3
L4 =
,
R
2 (M1 + M2 ) 2


R (M1 M2 )
3
L5 =
,
R
2 (M1 + M2 )
2

(23)

(24)

Stability and Application of the Lagrange Points


So much can be said about these results. In examining the stability of the mass placed at these equilibrium
points, done as a note later in the section, we nd that L4 and L5 are stabledue in part to the Coriolis
force which acts on a particle in motion; and L1 , L2 , and L3 are unstable. However, even though small
perturbations to a mass located at these Lagrange points would produce an unstable motion, the amplication
8

rate is small and it is far easier to locate a spacecraft at these points and apply an automatic control system
for station keeping than to locate a spacecraft at rest at an arbitrary point in space. So, even though
these point are points of unstable equilibrium, it requires only a small expenditure of fuel to keep spacecraft
at these points. Current space missions place objects in orbits about these Lagrange points for various
purposes, achieving very practical results.
Many space missions have used these points. For example the current plan is to locate the James Webb
optical telescope in an orbit about the Learthsun2 point where it will be shielded from direct sunlight by
the earth, and can be kept on station in a ecient manner.
In the natural world, groups of asteroids called the Trojan asteroidshave been observed to cluster at the
Sun-Jupiter L4 and L5 Lagrange points. Trojan asteroids have also been found at the L4 and L5 Lagrange
point of Mars and Neptune.
Note

Stability Analysis of the Lagrange Points

We study the stability of the various Lagrange points by expanding the equations of motion about each
equilibrium position, L1 to L5 . Setting
x = Lxi + xp

(25)

= Lyi + yp

(26)

= zp

(27)

where Lxi and Lyi are the x and y locations of the i th Lagrange (equilibrium) points and xp , yp , and
zp are small perturbations about these locations. We can for a very small price, include the z direction,
since the Lagrange points are located in the x, y plane, for which Lzi = 0. In studying the stability, we are
essentially deriving the equations for small oscillations about the equilibrium point; we intend to linearize
these equations. Oscillatory or decaying solutions imply stability; exponentially diverging solutions imply
instability. To obtain the equations governing small motions from equilibrium, we expand equation (12) for
small displacements about the various Lagrange points. Since the gradient of the potential is zero at these
equilibrium points, equation (12) in component form becomes
d2 xp
dyp
2
2
dt
dt
d 2 yp
dxp
+ 2
dt2
dt
d 2 zp
dt2

= Vxx xp Vxy yp Vxz zp

(28)

= Vxx xp Vxy yp Vxz zp

(29)

= Vxx xp Vxy yp Vxz zp

(30)

The various second derivatives of the pseudo potential V (x, y, z) are given in the table for the various
Lagrange points. Many of them are zero. This derivation allows the construction of the dierential equations
governing small displacement motion about the Lagrange points. From this we nd that L1 , L2 and L3 are
unstable, and L4 and L5 are stable. These equations provide the foundation for the analysis of station
9

Table 1: Stability Coecients for the Lagrange Points


point

V XX

V Y Y

L1

L2

92

L3

32
-

7
2
8

V ZZ

V XY

V XZ

42

L4

34 2

49 2

L5

34 2

49 2

V Y Z

3 4
3 2

3 4
3 2

keeping of a spacecraft at a Lagrange point. Thus, as previously mentioned, even though the motions of a
spacecraft located at one of the various co-linear Lagrange point would be unstable, the rate of amplication
is small and the Lagrange points are very attractive locations for various space missions.

10

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

S. Widnall
16.07 Dynamics
Fall 2009
Version 1.0

Lecture L19 - Vibration, Normal Modes, Natural Frequencies,

Instability

Vibration, Instability
An important class of problems in dynamics concerns the free vibrations of systems. (The concept of free
vibrations is important; this means that although an outside agent may have participated in causing an initial
displacement or velocityor both of the system, the outside agent plays no further role, and the subsequent
motion depends only upon the inherent properties of the system. This is in contrast to forced motion
in which the system is continually driven by an external force.) We shall consider only undamped systems
for which the total energy is conserved and for which the frequencies of oscillation are real. This forms the
basis of the approach to more complex studies for forced motion of damped systems. We saw in Lecture 13,
that the free vibration of a mass-spring system could be described as an oscillatory interchange between the
kinetic and potential energy, and that we could determine the natural frequency of oscillation by equating
the maximum value of these two quantities. (The natural frequency is the frequency at which the system
will oscillate unaected by outside forces. When we consider the oscillation of a pendulum, the gravitational
force is considered to be an inherent part of the system.) The general behavior of a mass-spring system can
be extended to elastic structures and systems experiencing gravitational forces, such as a pendulum. These
systems can be combined to produce complex results, even for one-degree of freedom systems.
We begin our discussion with the solution of a simple mass-spring system, recognizing that this is a model
for more complex systems as well.

In the gure, a) depicts the simple mass spring system: a mass M, sliding on a frictionless plane, restrained
by a spring of spring constant k such that a force F (x) = kx opposes the displacement x. (In a particular
problem, the linear dependence of the force on x may be an approximation for small x.) In order to get a
solution, the initial displacement and initial velocity must be specied. Common formulations are: x(0) = 0,
and

dx
dt (0)

= V0 (The mass responds to an initial impulse.); or x(0) = X0 and

dx
dt (0)

= 0 (The mass is given

an initial displacement.). The general formulation is some combination of these initial conditions. From
Newtons law, we obtain the governing dierential equation
m
with x(0) = X0 , and

dx
dt (0)

d2 x
= kx
dt2

(1)

= V0 .

The solution is of the general form, x(t) = Re(Aeit ), where, at this point in the analysis, both A and are
unknown. That is, we assume a solution in which both A and are unknown, and later when the solution
is found and boundary conditions are considered, we will end up taking the real part of the expression.
Depending upon whether A is purely real, purely imaginary, or some combination, we will in general get
oscillatory behavior involving sin s and cos s since eit = cos t + i sin t. For a system without damping,

2 = Real(K) (K being some combination of system parameters.), so that = K. For undamped


systems, these two values of , are redundant; only one need be taken.
To solve the dierential equation, equation(1) is rewritten
m

d2 x
+ kx = 0.
dt2

(2)

With the assumed form of solution, this becomes


2 mx(t) + kx(t) = 0 = x(t) ( 2 m + k)
Since x = x(t), for a valid solution, we require =

(3)

k
m.

This approach works as well for systems b) and c). These two systems are pendulums restrained by torsion
springs, which for small angles ( or ) produce a restoring torque proportional to the angular departure
from equilibrium. Consider system b). Its equilibrium position is = 0. The restoring torque from the
spring is Ts = . The restoring torque from gravity is Tg = mgLsin which for small angles becomes
Tg = mgL. Writing Newtons law in a form appropriate for pendular motion, we obtain
mL

d2
1
= (Tg + Ts )
dt2
L

(4)

We assume a form of solution, (t) = Ae(it) , and rewrite the equation as before, moving all terms to the
left-hand side.

1
mL 2 + (mgL + ) (t) = 0.
L

(5)

Therefore for a solution we require

mgL +
mL2
2

(6)

Examining this result, we see that the combination of the spring and gravity acts to increase the natural
frequency of the oscillation. Also if there is no spring, = 0, and the result becomes just the frequency of a

pendulum = Lg .
System c) is perhaps a bit more interesting. In this case, we use the small angle . We take the equilibrium
position of the spring to be = 0 so that the restoring torque due to the spring is again Ts = . But now
in this case, we are expanding the gravitational potential about the point = 0. Since this is an unstable
equilibrium point, this gives the restoring (it doesnt restore, it keeps going!) torque due to gravity as
Tg = mgLsin or for small , Tg = mgL.
Writing the governing equation for this case, we obtain

1
2
mL + (mgL + ) (t) = 0.
L

(7)

Therefore for a solution we require

mgL +
mL2

In this case, there is a critical value of for which = 0. On either side of this point, we have =

(8)

mgL
mL2 ,

which gives real for > mgL, and imag for < mgL. Since we have assumed (t) = eit ,
a real will produce oscillatory motion; an imaginary will produce exponentially diverging, or unstable,
motions. We say that the pendulum for less than the critical value, = mgL, is unstable.

Vibration of Multi-Degree of Freedom Systems


We begin our treatment of systems with multiple degrees of freedom, by considering a two degree of freedom
system. This system contains the essential features of multi-degree of freedom systems. Consider the two
two-mass, two-spring systems shown in the gure.

In this case, there are two independent variables, x1 (t) and x2 (t); their motion is not independent, but is
coupled by their attachments to the springs k1 , k2 and for system b), k3 . The sketch shows the forces Fi
acting on the masses as a result of the extension of the spring; these of are equal and opposite at the ends
of the springs. We consider both system a) and b). System b) is actually the simpler of the two systems
because of its inherent symmetry.
The governing equations can be written as
for system a)
d 2 x1
dt2
d2 x2
m2 2
dt
m1

= k1 x1 + k2 (x2 x1 )

(9)

= k2 (x2 x1 )

(10)

= k1 x1 + k2 (x2 x1 )

(11)

= k2 (x2 x1 ) k3 x2

(12)

for system b)
d 2 x1
dt2
d2 x2
m2 2
dt
m1

In both cases, as before, we assume a solution of the form x1 (t) = X1 eit and x2 (t) = X2 eit . However, as
we will see, in this case, we will obtain two possible values for 2 ; both will be real; we will take only the
positive value of itself. These will be the two vibration modes of this two degree of freedom system. These
results extend to N 2 s for an N degree of freedom system. Again, we will take only the positive value of
.
Consider rst system b). With the assumed form of solution, and rewriting all terms on the left-hand side,
we obtain
2 m1 X1 + k1 X1 k2 (X2 X1 )

=0

2 m2 X2 + k2 (X2 X1 ) + k3 X2

This equation can be written in matric form as


k1 + k2 k2
m 2

1
k2
k2 + k3
0
or

0
m2 2

k1 + k2 m1 2

k2

k2

k2 + k3 m2 2

X1
X2

(13)
0

X1
X2

(14)

0
0

0
0

(15)

(16)

This equation makes a very powerful statement. Since the right-hand side of both equations is zero, a
condition for a solution is that the determinant of the matrix equals zero. This will give an algebraic
equation with two solutions for : 1 and 2 . These are the natural frequencies of the two degree of
freedom system. In the general case, they are not equal; and both x1 and x2 participate in the oscillation
4

at each frequency i . Also, as in the single degree of freedom system, the actual values of x1 (t) and x2 (t)
are determined by initial conditions; in this case 4 initial conditions are required: x1 (0), x2 (0),
dx2
dt (0).

dx1
dt (0),and

The actual values of X1 and X2 are of less interest than the relationships between them and the

structure of the problem.

If the two masses are equal, a particularly simple form of a more general result follows from equation(16).

We consider this as an introduction to the more general case. The more general case will be considered

shortly.

k1 /m + k2 /m 2

k2 /m

k2 /m

k2 /m + k3 /m 2

X1
X2

0
0

(17)

We determine the two values of i (1 and 2 ) by setting the determinant equal to zero. We then substitute
each value of i in turn into the matrix equation and determine for each i the coecients X1i and X2i ;
1 = (X11 , X21 ) and
only their ratio can be determined. We write the coecients X1i and X2i as vectors, X
2 = (X12 , X22 ), where the subscript 1 refers to the mode associated with 1 , and 2 refers to the mode
X
associate with 2 .
It is a remarkable property of the solution to the governing equations that these vectors are orthogonal: the
1 X
2 = 0 (We will follow with an example to amplify and clarify this.)
dot product of X
Consider the simplest case of system b) with both masses equal to m and all springs of stiness k. In this
case we have

2k/m 2

k/m

k/m

2k/m 2

X1

.
(18)
0

Setting the determinant equal to zero gives two solutions for : 1 = k/m and 2 = 3k/m. The

X2

1 = (1, 1) and X
2 = (1, 1). This simple example gives great
components of the x1 and x2 motion are: X
i = (Xi1 , Xi2 )
physical insight to the more general problem. The natural frequency is i ; the components X
are called normal modes.

Normal Modes of Multi-Degree of Freedom Systems


1 = (1, 1) occurs at an oscillation
Examining the rst normal mode, we see an oscillation in which X

1 = (1, 1), the central spring does not deform, and the two masses oscillate,
frequency 1 = k/m. Since X

each on a single spring, thus giving a frequency = k/m.

2 = (1, 1); thus the masses move in


The second normal mode has a frequency = 3k/m, with X
opposite directions, and the frequency of oscillation is increased. It can be seen by inspection that the vector
1 and X
2 are orthogonal (their dot product is zero.)
X
If such a system was at rest, and an initial impulse was given to one of the masses, both modes would be
excited and a free oscillation would occur with each mode oscillating at its natural frequency.
The equation for general values of k1 , k2 and k3 can be written

k1 /m + k2 /m k2 /m
k2 /m

k2 /m + k3 /m

X1
X2

0
0

(19)

Characteristic Value Problem


This problem is called a characteristic-value or eigenvalue problem. Formulated in matrix notation it can be
written

A11

A12

A21

A22

X1
X2

0
0

(20)

The requirements on the simplest form of the characteristic value problem are that the matrix [A] is symmetric
(This will always be true for combinations of masses and springs.), and that the characteristic value
multiplies an identity matrix. (An identity matrix has all 1s on the diagonal and 0s o the diagonal; this
will be true if all the masses are equal; if not, a more general form must be used yielding analogous results.
This more general from will be considered shortly.)
For the form of the governing equation shown in equation(20), the characteristic value = 2 . The general
solution to this problem will yield a set of solutions for equal to the size of the matrix :(i.e a 4 4 matrix
i will be obtained. For this form of the characteristic value
will result in 4 s). For each i , a vector X
problem, the dot product between any two of these vectors is zero.
(Xi,1 , Xi,2 ).(Xj,1 , Xj,2 ) = 0

(21)

for i = j. These are the normal modes of the system, and the s are the natural frequencies. Any

numerical matrix methodsuch as MATLAB will yield both the i s (called the eigenvalues) and the Xi s,

called the eigenvectors for a particular matrix [A]. A similar result is obtained for the modes of vibration of

a continuous system such as a beam. The displacement of the various mode of vibration of a uniform beam

are orthogonal.

i ,

The general solution for the motion of the masses is then given by an expansion in the normal modes, X

x1 (t)
X
X

= A1 ei1 t 11 + A2 ei2 t 21 ,
(22)
x2 (t)
X12
X22
1 = (X11 , X21 ) and X
2 = (X12 , X22 ) are the eigenvectors or normal modes from the solution of the
where X
characteristic-value problem, obtained by hand or numerically, and i is the natural frequency of that mode.
i and X
j is zero unless
Again, it is a remarkable and extremely useful property that the dot product of X
i = j. Good form would suggest that we normalize each Xi so that the magnitude of Xi Xi equals 1, but
i X
i , the vector-magnitude-squared, for later use.
we usually dont and therefore need to dene Ci = X

Expansion in Normal Modes; Satisfaction of Initial Conditions


The general form of solution is given by equation (22). All that remains is to determine the coecients A1
and A2 . This is done by satisfying the initial conditions on displacement and velocity. In the general case,
since eit = 1 at t = 0, the initial displacement can be written

x1 (0)
X11
X21

= A1
+ A2
.
x2 (0)
X12
X22
and for an initial velocity, since ieit = i at t = 0, the initial velocity can be written

v1 (0)
X11
X21

= iA1
+ iA2
.
v2 (0)
X12
X22

(23)

(24)

It should be noted that in general A is complex; the real part relates to the initial displacement; the imaginary
part to the initial velocity. If we consider Ai to be real, we are automatically assuming no initial velocity.
Case 1: initial displacement non-zero; initial velocity zero
We rst consider the case of an initial condition on the displacement, specically x1 (0) = x10 and x2 (0) = x20 ,
0 = (x1 (0), x2 (0)). To complete
with v1 (0) = 0 and v2 (0) = 0. We dene the initial-condition vector as X
the solution, we need to obtain the values of A1 and A2 from equation(23). This is done by taking the dot
1 . Finally we get our payo for all our hard work.
product of both sides of equation(23) with the rst mode X
1 with X
2 is zero; the dot product of X
1 with X
1 is C1 ; and the dot product of
Since the dot product of X
0 with X
1 is some G1 , we obtain
X
0 X
1 /C1 = G1
A1 = X
C1

(25)

0 with X
2 (which equals some G2 ), we obtain
and taking the dot product of X
0 X
2 /C2 = G2
A2 = X
C2

(26)

With A1 and A2 determined, we have a complete solution to the problem.


Case 2: initial displacement zero; initial velocity non-zero
We now consider the case of an initial condition on the velocity, specically x 1 (0) = v10 and x 2 (0) = v20 , with
x1 (0) = 0 and x2 (0) = 0. We dene the initial-condition vector as V0 = (v10 , v20 ). We use the coecient Bi
to dene the solution for this case. The solution is again written as an expansion in normal modes oscillating
at their natural frequency i of amplitude Bi , which is unknown at this point.

x1 (t)
X
X
11
21

= B1 eit
+ B2 eit
.
x2 (t)
X12
X22

(27)

As previously noted, B is complex; the real part relates to the initial displacement; the imaginary part to
the initial velocity. If Bi is imaginary, there is no initial displacement. The velocity at t = 0 is given by

v1 (0)
X11
X21

= iB1
+ iB2
.
(28)
v2 (0)
X12
X22
To complete the solution, we need to obtain the values of B1 and B2 from equation(28). This is again done
1 . Since the dot product of X
1 with X
2 is
by taking the dot product of equation(28) with the rst mode X
1 with X
1 is C1 , we obtain
zero, and the dot product of X
B1 =

Q1 i
C1

(29)

1 and taking the dot product of X


0 with X
2 we obtain
where Q1 = V0 X
B2 =

Q2 i
C2

(30)

2 and X
2 X
2 C2 =. The fact that Bi is purely imaginary conrms our earlier observation
where Q2 = V0 X
that real coecients imply a non-zero initial displacement while purely imaginary coecients imply a non
zero initial velocity. A purely imaginary Bi simply implies that the displacement has a sin(t) behavior in
contrast to a cos(t) behavior, since the real part of eit = cos(t), while the real part of ieit = sin(t).
and x2 (0)
would be a linear combination of
A solution for general initial conditions on x1 (0), x2 (0), x1 (0)
these solutions.

Solution for Unequal Masses


If the masses, mi , are not equal we must use a more general form of the eigenvalue problem. Returning to
Equation (19) for equal masses.
For the case of equal masses, from Equation (19), this can be written

k1 /m + k2 /m k2 /m
X1
1

= 2
k2 /m
k2 /m + k3 /m
X2
0
For unequal masses, we rewrite the equation as

k1 + k2 k2
X1
m

= 2 1
k2
k2 + k3
X2
0

0
m2

0
1

X1
X2

(31)

(32)

This equation is in the form of the generalized characteristic or eigenvalue problem. (See Hildebrand;
Methods of Applied Mathematics.)

a11 a12
b

1
a21 a22
0

0
b2

X1
X2

0
0

(33)

where equation (20) has been rewritten as (33) in the extended formulation ([A] [B])(X) = 0. Following
Hildebrand we note that both [A] and [B] are symmetric matrices. Moreover, [B] is a diagonal matrix. Fol
lowing the solution of the generalized characteristic or eigenvalue problem, by numerical or other technique,
the resulting orthogonality condition between the normal modes is modied as

Xj,1
m1 0

=0
(Xi,1 , Xi,2 )
0
m2
Xj,2

(34)

for i =
j. Thus the vectors of displacement for the normal modes of vibration must be multiplied by the
mass distribution to result in orthogonality. A similar result is obtained for the vibration of a continuous
system, such as a beam with non-uniform mass distribution. The process of expanding the solution in term
of normal modes goes through as before with the modication of the normality condition.
In more general congurations, ([A] [B])(X) = 0, the matrix [B] may not be diagonal, that is

b11 b12

[B] =
b21 b22
As long as [B] is symmetric, the orthogonalization goes through as

b11 b12
Xj,1

=0
(Xi,1 , Xi,2 )
b21 b22
Xj,2

(35)

(36)

Observations
The process we have outlined for nding the solution to the initial value problem to a multi-degree of
freedom system, outlined from equation(20) on, works for system with degrees of freedom from 2 to 20,000
and beyond. This approach is of fundamental importance in analyzing vibrations in a wide variety of systems.
The expansion in normal modes is also useful in more complex problems such as forced motions at frequencies
other than i .

References
[1] Hildebrand: Methods of Applied Mathematics; for a discussion of the characteristic value problem,
matrices and vectors.

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

S. Widnall
16.07 Dynamics
Fall 2009
Version 3.0

Lecture L20 - Energy Methods: Lagranges Equations


The motion of particles and rigid bodies is governed by Newtons law. In this section, we will derive an
alternate approach, placing Newtons law into a form particularly convenient for multiple degree of freedom
systems or systems in complex coordinate systems. This approach results in a set of equations called
Lagranges equations. They are the beginning of a complex, more mathematical approach to mechanics
called analytical dynamics. In this course we will only deal with this method at an elementary level. Even at
this simplied level, it is clear that considerable simplication occurs in deriving the equations of motion for
complex systems. These two approachesNewtons Law and Lagranges Equationsare totally compatible.
No new physical laws result for one approach vs. the other. Many have argued that Lagranges Equations,
based upon conservation of energy, are a more fundamental statement of the laws governing the motion of
particles and rigid bodies. We shall not enter into this debate.

Derivation of Lagranges Equations in Cartesian Coordinates


We begin by considering the conservation equations for a large number (N) of particles in a conservative
force eld using cartesian coordinates of position xi . For this system, we write the total kinetic energy as

T =

1
mi x 2i .
2
n=1

(1)

where M is the number of degrees of freedom of the system.

For particles traveling only in one direction, only one xi is required to dene the position of each particle,

so that the number of degrees of freedom M = N . For particles traveling in three dimensions, each particle

requires 3 xi coordinates, so that M = 3 N .

The momentum of a given particle in a given direction can be obtained by dierentiating this expression

with respect to the appropriate xi coordinate. This gives the momentum pi for this particular particle in

this coordinate direction.

T
= pi
x i

(2)

The time derivative of the momentum is


d
dt

T
x i

= mi x
i

(3)

For a conservative force eld, the force on a particle is given by the derivative of the potential at the particle
position in the desired direction.

Fi =

V
xi

(4)

From Newtons law we have


Fi =

dpi
dt

(5)

Equating like terms from our manipulations on kinetic energy and the potential of a conservative force eld,
we write
d
dt

T
x i

V
xi

(6)

Now we make use of the fact that


T
=0
xi

(7)

V
=0
x i

(8)

and

Using these results, we can rewrite Equation (6) as

d (T V )
(T V )

=0
dt
x i
xi

(9)

We now dene L = T V : L is called the Lagrangian. Equation (9) takes the nal form: Lagranges
equations in cartesian coordinates.
d
dt

L
x i

L
=0
xi

(10)

where i is taken over all of the degrees of freedom of the system. Before moving on to more general coordinate
systems, we will look at the application of Equation(10) to some simple systems.

Mass-spring System

We rst consider a simple mass spring system. This is a one degree of freedom system, with one xi . Its
kinetic energy is T = 1/2mx 2 ; its potential is V = 1/2kx2 ; its Lagrangian is L = 1/2mx 2 1/2kx2 . Applying
Equation (10) to the Lagrangian of this simple system, we obtain the familiar dierential equation for the
mass-spring oscillator.
m

d2 x
+ kx = 0
dt2

(11)

Clearly, we would not go through a process of such complexity to derive this simple equation. However, lets
consider a more complex system, governed by the same laws.

This is a 2 degree of freedom system, governed by 2 dierential equations. The number of springs for
this conguration is 3. These governing equations could be obtained by applying Newtons Law to the
force balance that exists at each mass due to the deection of the springs as was done in Lecture 19. The
deection of springs 1 and 3 are inuenced by the boundary condition at either end of the slot; in this case
the deection is zero.
The governing equations can also be obtained by direct application of Lagranges Equation. This approach
is quite straightforward. The expression for kinetic energy is
T =

1
mi x 2i ;
2
n=1

(12)

the expression for the potential is


V = 1/2k1 x21 + 1/2k2 (x2 x1 )2 + 1/2k3 x22
Applying Lagranges equation to T V = L

d L

dt x 1

d L

dt x 2

L
=0
x1
L
=0
x2

(13)

(14)
(15)

we obtain the governing equations as

d2 x1
dt2
d2 x2
m2 2
dt
m1

= k1 x1 + k2 (x2 x1 )

(16)

= k2 (x2 x1 ) k3 x2

(17)

Clearly, for multi degree of freedom systems, this approach has advantages over the force balancing approach
using Newtons law.

Extension to General Coordinate Systems


A signicant advantage of the Lagrangian approach to developing equations of motion for complex systems
comes as we leave the cartesian xi coordinate system and move into a general coordinate system. An
3

example would be polar coordinates where for a two-dimensional position of a mass particle, x1 and x2
could be given by r and . A two-degree of freedom system remains two-degree so that the number of
coordinate variables required remains two. r and and their counterparts in other coordinate systems
will be referred to as generalized coordinates. We introduce quite general notation for the relationship
between the n cartesian variables of position xi and their description in generalized coordinates. (For
some systems, the number of generalized coordinates is larger than the number of degrees of freedom and
this is accounted for by introducing constraints on the system. This is an important part of the discussion
of the Lagrange formulation. We shall not however develop these relations but will work directly with the
number of variables equal to the number of degrees of freedom of the system.) We express the cartesian
variable xi using generalized coordinates qj . (Polar coordinates r, would be an example.)

xi = xi (q1 , ..qj , ...qn ).

(18)

In the general case, each xi could be dependent upon every qj .

What is remarkable about the Lagrange formulation, is that (10) holds in a general coordinate system with

xi replaced by qi .
d
dt

L
qi

=0
qi

(19)

Before showing how this result can be derived from Newtons Law, we show two applications in polar
coordinates to demonstrate the power of the approach.

Simple Pendulum by Lagranges Equations


We rst apply Lagranges equation to derive the equations of motion of a simple pendulum in polar coor
dinates. This is a one degree of freedom system. However, it is convenient for later analysis of the double
pendulum, to begin by describing the position of the mass point m1 with cartesian coordinates x1 and y1
and then express the Lagrangian in the polar angle 1 . Referring to a) in the gure below we have
x1 = h1 sin 1

(20)

y1 = h1 cos 1

(21)

1
1
m1 (x 21 + y 12 ) = m1 h21 12
2
2

(22)

so that the kinetic energy is


T =
The potential energy is
V = m1 gy1 = m1 gh1 cos

(23)

The Lagrangian is
L=T V =

1
m1 h21 12 + m1 gh1 cos 1
2

(24)

Applying (10) with q1 = 1 , we obtain the dierential equation governing the motion.
m1 h21 1 + m1 gh1 sin 1 = 0

(25)

Again, for such a simple system, we would typically not go through this formalism to obtain this result.
However, this framework will enable us to derive the equations of motion for the more complex systems such
as the double pendulum shown in b).

Double Pendulum by Lagranges Equations


Consider the double pendulum shown in b) consisting of two rods of length h1 and h2 with mass points m1
and m2 hung from a pivot. This systems has two degrees of freedom: 1 and 2 .

To apply Lagranges equations, we determine expressions for the kinetic energy and the potential as the
system moves in angular displacement through the independent angles 1 and 2 . From the geometry we
have

The kinetic energy is T =

1
2 m1

x1 = h1 sin 1

(26)

y1 = h1 cos 1

(27)

x2 = h1 sin 1 + h2 sin 2

(28)

y2 = h1 cos 1 h2 cos 2

(29)

x 21 + y 12 + 12 m2 x 22 + y 22 . Expressed in variables 1 and 2 , the kinetic

energy of the system is


T =

1
1
m1 h21 12 + m2 h21 12 + h22 22 + 2h1 h2 1 2 cos(1 2 )
2
2

(30)

The potential energy of the system is


V = m1 gy1 + m2 gy2 = (m1 + m2 )gh1 cos 1 m2 gh2 cos 2 .

(31)

The Lagrangian is then


L = T V =

1
1
(m1 +m2 )h21 12 + m2 h22 22 +m2 h1 h2 1 2 cos(1 2 )+(m1 +m2 )gh1 cos 1 +m2 gh2 cos 2 (32)
2
2

Since the generalized coordinates are now 1 and 2 , Lagranges equation becomes

d L
L

=0

dt i
i

(33)

for both i = 1 and i = 2.

d L

dt 1

d L

dt 2

L
=0
1
L
=0
2

Working out the details, we have

d L
= (m1 + m2 )h21 1 + m2 h1 h2 2 cos(1 2 ) m2 h1 h2 2 sin(1 2 )(1 2 )
dt 1
L
= h1 g(m1 + m2 ) sin(1 ) m2 h1 h2 1 2 sin(1 2 )
1

d L
= m2 h22 2 + m2 h1 h2 1 cos(1 2 ) m2 h1 h2 1 sin(1 2 )(1 2 )
dt 2
L
= h2 gm2 sin(2 ) + m2 h1 h2 1 2 sin(1 2 )
2

(34)
(35)

(36)
(37)
(38)
(39)

From Equation(35-39), we obtain the nal form of the governing equations for the double pendulum
(m1 + m2 )h1 1 + m2 h2 2 cos(1 2 ) + m2 h2 22 sin(1 2 ) + g(m1 + m2 ) sin 1 = 0

(40)

m2 h2 2 + m2 h1 1 cos(1 2 ) m2 h1 12 sin(1 2 ) + m2 g sin 2 = 0

(41)

The double pendulum is a system of great interest, displaying conventional linear multi degree of freedom sys
tem behavior for small 1 and 2 , but displaying chaotic behavior for large . A chaotic system is a determin
istic system that exhibits great sensitivity to the initial conditions: the buttery eect. A simulation of the
motion of a double pendulum is available on http://scienceworld.wolfram.com/physics/DoublePendulum.html.
For a particular choice of initial conditions, the position of m2 with time is shown in the gure.

Derivation of Lagranges Equation for General Coordinate Systems


We now follow the earlier procedure we used to derive Lagranges equation from Newtons law but using
generalized coordinates instead of cartesian coordinates. (See additional reading: Slater and Frank and/or
Marion and Thorton.) The cartesian variable are expressed in terms of the generalized coordinates as

xi = xi (q1 , ..qj , ...qn ).

(42)

To obtain the velocity x i , we take the time derivative of (42) applying the chain rule.

x i =

xi
j=1

qj

qj .

(43)

Taking the partial derivative of (43) with respect to qj we obtain a relation between these two derivatives,
x i
xi
=
,
qj
qj

(44)

a result which we shall need shortly. In deriving equation (44), we take advantage of the fact that since the
generalized coordinates qi are independent,

qi
qj

= 0 for i =
j.

Following the approach leading to Equations (2-10), we dene a generalized momentum as

pi =

T
qi

(45)

1
mj x 2j as given by Equation (1).
2
n=1
We are now ready to express Newtons law in the generalized coordinates qi that we have introduced.

with T =

For the generalized momentum we have


n

pi =

T
x j
xj
=
mj x j
=
mj x j
,
qi
q

qi
i
j=1
j=1

(46)

where we have made use of (44).

For a conservative force eld, the work done during a displacement dxi is given by

dW =

Fi dxi

(47)

i=1

or expressed in the generalized coordinates qj


dW =

Fi dxi =

N
N

i=1

Fi

j=1 i=1

xi
dqj
qj

(48)

We identify
N

Fi

i=1

xi
dqj
qj

(49)

as the work done by a displacement through dqj and dene a generalized force Qj as
Qj =

xi
qj

(50)

Qj dqj

(51)

Fi

i=1

so that the work is expressed as.


dW =

j=1

For a conservative system, the work done by a small displacement dqj is


dW = dV =

V
dqj ,
qj

(52)

where V is the potential function expressed in the coordinate system of the generalized coordinates and

V
qj

is the change in potential due to a change in the generalized coordinate qj . So that the generalized force
is
Qj =

V
qj

(53)

in analogy with (4). We now examine the time derivative of the generalized momentum, pi .
N

dpi
d T
xj
d xj
= (
)=
(mj x
j
+ mj x j
).
dt
dt qi
q
dt
qi
i
j=1

(54)

Since xj /qi is a function of the q s which are functions of time, we have by the chain rule
N

2 xj
d xj
(
)=
qk
dt qi
qi qk

(55)

k=1

From Newtons law, mj x


j = Fj , so that the rst term in (54) is given by
N

j=1

mj x
j

xj
= Qi
qi
8

(56)

Now let us consider the second term of (54). Consider the expression for T /qi and use (43, 44).
N
T
x j
xi
= mj x j
= mj x j
qk
qi
qi
qi
qk

(57)

k=1

This is precisely the second term in equation (54). We have now identied simpler forms for the two
expressions in the equation for the time derivative of the generalized momentum. We may now write
dpi
d
=
dt
dt

T
qi

= Qi +

T
V
T
=
+
qi
qi
qi

(58)

Since for a conservative system, the potential is independent of the velocities, we can place this equation
into nal form by dening the Lagrangian as L = T V to obtain the nal form of Lagranges equation as
d
dt

L
qi

L
=0
qi

(59)

in agreement with (19). Equation (59) is Lagranges Equations in generalized coordinates. In our previous
example, we applied this equation to simple and double pendulums in polar coordinates using q1 = 1 and
q2 = 2 . What is signicant about this equation, adding to its power, is that each i equation contains only
derivatives with respect to that qi and qi .
ADDITIONAL READING
Slater and Frank, Mechanics, Chapter IV
Marion and Thorton, Chapter 7

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

J. Peraire, S. Widnall
16.07 Dynamics
Fall 2008
Version 2.0

Lecture L21 - 2D Rigid Body Dynamics

Introduction
In lecture 11, we derived conservation laws for angular momentum of a system of particles, both about the
center of mass, point G, and about a xed (or at least non-accelerating) point O. We then extended this
derivation to the motion of a rigid body in two-dimensional plane motion including both translation and
rotation. We obtained statements about the conservation of angular momentum about both a xed point
and about the center of mass. Both are powerful statements. However, each has its own sperate requirements
for application. In the case of motion about a xed point, the point must have zero acceleration. Thus
the instantaneous center of rotation, for example the point of contact of a cylinder rolling on a plane, cannot
be used as the origin of our coordinates. For motion about the center of mass, no such restriction applies
and we may obtain the statement of conservation of angular momentum about the center of mass even if
this point is accelerating.

Kinematics of Two-Dimensional Rigid Body Motion


Even though a rigid body is composed of an innite number of particles, the motion of these particles is
constrained to be such that the body remains a rigid body during the motion. In particular, the only degrees
of freedom of a 2D rigid body are translation and rotation.

Parallel Axes
Consider a 2D rigid body which is rotating with angular velocity about point O , and, simultaneously,
point O is moving relative to a xed reference frame x and y with origin O.

In order to determine the motion of a point P in the body, we consider a second set of axes x y , always
parallel to xy, with origin at O , and write,
rP

= r O + r P

(1)

vP

= v O + (v P )O

(2)

aP

= aO + (aP )O .

(3)

Here, r P , v P and aP are the position, velocity and acceleration vectors of point P , as observed by O; r O
is the position vector of point O ; and r P , (v P )O and (aP )O are the position, velocity and acceleration
vectors of point P , as observed by O . Relative to point O , all the points in the body describe a circular
orbit (rP = constant), and hence we can easily calculate the velocity,
(vP )O = rP = r ,
or, in vector form,
(v P )O = r P ,
where is the angular velocity vector. The acceleration has a circumferential and a radial component,
((aP )O ) = rP = rP ,

((aP )O )r = rP 2 = rP 2 .

Noting that and are perpendicular to the plane of motion (i.e. can change magnitude but not
direction), we can write an expression for the acceleration vector as,
(aP )O = r P + ( r P ) .
Recall here that for any three vectors A, B and C, we have A (B C) = (A C)B (A B)C. Therefore

( r P ) = ( r P
) 2 r P = 2 r P . Finally, equations 2 and 3 become,

vP

= v O + r P

(4)

aP

= aO + r P + ( r P ) .

(5)

Body Axes
An alternative description can be obtained using body axes. Now, let x y be a set of axes which are rigidly
attached to the body and have the origin at point O .

Then, the motion of an arbitrary point P can be expressed in terms of the general expressions for relative
motion. Recall that,
rP

= r O + r P

(6)

vP

= v O + (v P )O + r P

(7)

aP

= aO + (aP )O + 2 (v P )O + r P + ( r P ) .

(8)

Here, r P , v P and aP are the position, velocity and acceleration vectors of point P as observed by O; r O is
the position vector of point O ; r P , (v P )O and (aP )O are the position, velocity and acceleration vectors of
= are the body angular velocity and acceleration.
point P as observed by O ; and = and
,
Since we only consider 2D motions, the angular velocity vector, , and the angular acceleration vector,
do not change direction. Furthermore, because the body is rigid, the relative velocity (vP )O and acceleration
(aP )O of any point in the body, as observed by the body axes, is zero. Thus, equations 7 and 8 simplify to,
vP

= v O + r P

aP

= aO + r P + ( r P ) ,

(9)
(10)

which are identical to equations 4 and 5, as expected. Note that their vector forms are equal. If at t=0,
the frame x , y (and eventually z ) are instantaneously aligned with the frame x, y, the components of the
vectors are qual. If not, then a coordinate transformation is required.

Invariance of and =
The angular velocity, , and the angular acceleration, = , are invariant with respect to the choice of
the reference point O . In other words, this means that an observer using parallel axes situated anywhere in
the rigid body will observe all the other points of the body turning around, in circular paths, with the same
angular velocity and acceleration. Mathematically, this can be seen by considering an arbitrary point in the
body O and writing,

r P = r O
+ r P .

Substituting into equations 6, 9 and 10, we obtain,


rP

= r O + r O + r P

(11)

vP

= v O + r O + r P

(12)

aP

= aO + r O + r P + ( r O ) + ( r P ) .

(13)

From equations 6, 9 and 10, replacing P with O , we have that r O = r O + r O , v O = v O + r O ,


and aO = aO + r O + ( r O ). Therefore, we can write,
rP

= r O + r P

(14)

vP

= v O + r P

(15)

aP

= aO + r P + ( r P ) .

(16)

These equations show that if the velocity and acceleration of point P are referred to point O rather than
point O , then r P =
r P , v O = v O , and aO = aO , although the angular velocity and acceleration vectors,
and , remain unchanged.

Instantaneous Center of Rotation


We have established that the motion of a solid body can be described by giving the position, velocity and
acceleration of any point in the body, plus the angular velocity and acceleration of the body. It is clear
that if we could nd a point, C, in the body for which the instantaneous velocity is zero, then the velocity
of the body at that particular instant would consist only of a rotation of the body about that point (no
translation). If we know the angular velocity of the body, , and the velocity of, say, point O , then we
could determine the location of a point, C, where the velocity is zero. From equation 9, we have,
0 = v O + r C .
Point C is called the instantaneous center of rotation. Multiplying through by , we have v O =
( r C ), and, re-arranging terms, we obtain,
r C =

1
( v O ) ,
2

which shows that r C and v O are perpendicular, as we would expect if there is only rotation about C.
Alternatively, if we know the velocity at two points of the body, P and P , then the location of point C
can be determined geometrically as the intersection of the lines which go through points P and P and are
perpendicular to v P and v P . From the above expression, we see that when the angular velocity, , is very
small, the center of rotation is very far away, and, in particular, when it is zero (i.e. a pure translation),
the center of rotation is at innity. Although the center of rotation is a useful concept, if the pointO is

accelerating, it cannot be used as the origin in the application of the principle of conservation of angular
momentum.
Example

Rolling Cylinder

Consider a cylinder rolling on a at surface, without sliding, with angular velocity and angular acceleration
. We want to determine the velocity and acceleration of point P on the cylinder. In order to illustrate the
various procedures described, we will consider three dierent approaches.

Direct Method :

Here, we nd an expression for the position of P as a function of time. Then, the velocity and

acceleration are obtained by simple dierentiation. Since there is no sliding, we have,

v O = R i,

aO = R i,

and,
rP

= r O + R cos i + R sin j .

Therefore,
v P = r P

= v O R sin i + R cos j
=

aP = r P

R(1 + sin ) i + R cos j

= aO R( sin + cos ) i + R( cos sin ) j


=

[R(1 + sin ) 2 R cos ] i + [R cos 2 R sin ] j

Relative motion with respect to C :


Here, we use expressions 4 and 5, or 9 and 10, with O replaced by C.
vP

= r CP

(17)

aP

= aC + r CP + ( r CP ) .

(18)

In the above expressions, we have already used the fact that v C = 0. Now,
r CP = R cos i + R(1 + sin ) j ,

= k ,

and,
v P = R(1 + sin ) i + R cos j .
The calculation of aP , in this case, requires knowing aC . In the no sliding case, aC can be shown to be
equal to R 2 j, i.e., it only has a vertical component. With this, after some algebra, we obtain,
aP = [R(1 + sin ) 2 R cos ] i + [R cos 2 R sin ] j .

Example

Sliding bar

Consider a bar leaning against the wall and slipping downward. It is clear that while the bar is in contact
with the wall and the oor, the velocity at point P will be in the vertical direction, whereas the velocity
at point P will be in the horizontal direction. Therefore, drawing the perpendicular lines to v P and v P
through points P and P , we can determine the instantaneous center of rotation C.

It should be noted that, for a general motion, the location of the center of rotation will change in time.
The path described by the instantaneous center of rotation is called the space centrode, and the locus of the
positions of the instantaneous centers on the body is called the body centrode. At a given instant, the space
centrode and the body centrode curves are tangent. The tangency point is precisely the instantaneous center
of rotation, C. Therefore at this instance, the point C is common to both curves. It is not dicult to show
that, for the above example, the space and body centrodes are circular arcs, assuming that the points P and
P remain in contact with the walls at all times.

From this example, it should be clear that although we think about the instantaneous center of rotation as
a point attached to the body, it need not be a material point. In fact, it can be a point outside the body.
It is also possible to consider the instantaneous center of acceleration as the point at which the instantaneous
acceleration is zero.

Review: Conservation of Angular Momentum for 2D Rigid Body


In Lecture 11, we showed that the equations describing the general motion of a rigid body follow from the
conservation laws for systems of particles. Since the general motion of a 2D rigid body can be determined
by three parameters (e.g. x and y coordinates of position, and a rotation angle ), we will need to supply
three equations. Conservation of linear momentum yields one vector equation, or two scalar equations. The
additional condition is conservation of angular momentum. In Lecture 11, We saw that there are several
ways to express conservation of angular momentum. In principle, they are all equivalent, but, depending on
the problem situation, the use of a particular form may greatly simplify the problem. The best choices for
the origin of coordinates are: 1) the center of mass G; 2) a xed point O.

Conservation of Angular Momentum about the Center of Mass


When considering a 2D rigid body, the velocity of any point relative to G consists of a pure rotation and,
therefore, the conservation law for angular momentum about the center of mass, G is
HG =

(r i mi ( r i )) =

i=1

i=1

mi ri 2 =

r2 dm

(19)

For a continuous body, the sum over the mass points is replaced by an integral.

r2 dm is dened as the

mass moment of inertia IG about the center of mass.


IG = M G ,
7

(20)

where IG =

r2 dm, = and MG is the total moment about G due to external forces and external

moments. Although equation (7) is a vector equation, and M G are always perpendicular to the plane
of motion, and, therefore, equation (7) only yields one scalar equation. The moment of inertia, IG , can
be interpreted as a measure of the bodys resistance to changing its angular velocity as a result of applied
external moments. The moment of inertia, IG , is a scalar quantity. It is a property of the solid which
indicates the way in which the mass of the solid is distributed relative to the center of mass. For example, if
most of the mass is far away from the center of mass, ri will be large, resulting in a large moment of inertia.
The dimensions of the moment of inertia are [M ][L2 ].

Conservation of Angular Momentum about a xed point O


If the xed point O is chosen as the origin, a similar result is obtained. Since for a 2D rigid body the velocity
of any point in the coordinate system xed at the point O is
vi = ri ,
conservation of angular momentum gives

IO = M O ,
where IO =

(21)

r2 dm, = and MO is the total applied moment due to external forces and moments

(torques). Also it is important to point out that both the angular velocity and the angular acceleration

are the same for any point on a rigid body: G = O , G = O .

Most textbooks on dynamics have tables of moments of inertia for various common shapes: cylinders, bars,

plates. See Meriam and Kraige, Engineering Mechanics, DYNAMICS (Appendix B) for more examples.

Radius of Gyration
It is common to report the moment of inertia of a rigid body in terms of the radius of gyration, k. This is
dened as

I
,
m
and can be interpreted as the root-mean-square of the mass element distances from the axis of rotation.
k=

Since the moment of inertia depends upon the choice of axis, the radius of gyration also depends upon the
choice of axis. Thus we write

IG
,
m
for the radius of gyration about the center of mass, and

IO
kO =
,
m
kG =

for the radius of gyration about the xed point O.


8

Parallel Axis Theorem


We will often need to nd the moment of inertia with respect to a point other than the center of mass. For
instance, the moment of inertia with respect to a given point, O, is dened as

IO =
r2 dm .
m

Assuming that O is a xed point, H O = IO . If we know IG , then the moment of inertia with respect
to point O, can be computed easily using the parallel axis theorem. Given the relations r2 = r r and
r = r G + r ., we can then write,

2
2

2
2
IO =
r dm =
(rG + 2r G r + r ) dm = rG
m

since

dm + 2r G

2
r2 dm = mrG
+ IG ,

r dm +
m

r dm = 0.

From this expression, it also follows that the moment of inertia with respect to an arbitrary point is mini
mum when the point coincides with G. Hence, the minimum value for the moment of inertia is IG .

Summary: Governing Equations


Now that we have reviewed the equation governing conservation of angular momentum for a 2D rigid body

in planar motion about both the center of mass and about a xed point O , we can restate the governing

equations for this three degree of freedom system.

The conservation of linear momentum yields the vector equation,

maG = F ,

(22)

where m is the body mass, aG is the acceleration of the center of mass, and F is the sum of the external
forces acting on the body.
Conservation of angular momentum about the center of mass requires
G = M G = IG = IG
H

(23)

where IG is the moment of inertia about the center of mass; is the angular velocity, whose vector direction

is perpendicular to the x, y coordinate system; and a is the angular acceleration.

A body xed at a point O is a single degree of freedom system. Therefore, only one equation is required,

conservation of angular momentum about the point O.

O = M O = IO = IO
H

Kinetic Energy for a 2D Rigid Body


We start by recalling the kinetic energy expression for a system of particles derived in lecture L11,
9

(24)

T =

1
i=1

mi (v G + r i ) (v G + r i ) =

1
1
2
mvG
+
2
2

r2 dm =

1
1
2
mvG
+ 2 IG .
2
2

(25)

where n is the total number of particles, mi denotes the mass of particle i, and r i is the position vector of
n
particle i with respect to the center of mass, G. Also, m = i=1 mi is the total mass of the system, and
v G is the velocity of the center of mass. The above expression states that the kinetic energy of a system of
particles equals the kinetic energy of a particle of mass m moving with the velocity of the center of mass,
plus the kinetic energy due to the motion of the particles relative to the center of mass, G.
2
When the body is rotating about a xed point O, we can write IO = IG + mrG
and

T =

1
1
1
2
2
mvG
+ (IO mrG
) 2 = IO 2 ,
2
2
2

since vG = rG .

The above expression is also applicable in the more general case when there is no xed point in the motion,

provided that O is replaced by the instantaneous center of rotation. Thus, in general,

T =

1
IC 2 .
2

We shall see that, when the instantaneous center of rotation is known, the use of the above expression does
simplify the algebra considerably.

Work
External Forces
Since the body is rigid and the internal forces act in equal and opposite directions, only the external forces
applied to the rigid body are capable of doing any work. Thus, the total work done on the body will be
n

i=1

(Wi )12 =

i=1

(r i )2

F i dr ,

(r i )1

where F i is the sum of all the external forces acting on particle i.

10

Work done by couples


If the sum of the external forces acting on the rigid body is zero, it is still possible to have non-zero work.
Consider, for instance, a moment M = F a acting on a rigid body. If the body undergoes a pure translation,
it is clear that all the points in the body experience the same displacement, and, hence, the total work done
by a couple is zero. On the other hand, if the body experiences a rotation d, then the work done by the
couple is
dW = F

a
a
d + F d = F ad = M d .
2
2

If M is constant, the work is simply W12 = M (2 1 ). In other words, the couples do work which results
in the kinetic energy of rotation.

Conservative Forces
When the forces can be derived from a potential energy function, V , we say the forces are conservative. In
such cases, we have that F = V , and the work and energy relation in equation ?? takes a particularly
simple form. Recall that a necessary, but not sucient, condition for a force to be conservative is that it
must be a function of position only, i.e. F (r) and V (r). Common examples of conservative forces are gravity
(a constant force independent of the height), gravitational attraction between two bodies (a force inversely
proportional to the squared distance between the bodies), and the force of a perfectly elastic spring.
The work done by a conservative force between position r 1 and r 2 is
r2
r
W12 =
F dr = [V ]r21 = V (r 1 ) V (r 2 ) = V1 V2 .
r1

NC
Thus, if we call W12
the work done by all the external forces which are non conservative, we can write the

general expression,
NC
T1 + V1 + W12
= T2 + V2 .

11

Of course, if all the forces that do work are conservative, we obtain conservation of total energy, which can
be expressed as,
T + V = constant .

Gravity Potential for a Rigid Body


In this case, the potential Vi associated with particle i is simply Vi = mi gzi , where zi is the height of particle
i above some reference height. The force acting on particle i will then be F i = Vi . The work done on
the whole body will be
n

i=1

r 2i

r 1i

F i dr i =

((Vi )1 (Vi )2 ) =

i=1

mi g((zi )1 (zi )2 = V1 V2 ,

i=1

where the gravity potential for the rigid body is simply,


V =

mi gzi = mgzG ,

i=1

where zG is the z coordinate of the center of mass. Its obvious but worth noting that because the gravitational
potential is taken about the center of mass, the inertia plays no role in determining the gravitational potential.

ADDITIONAL READING
J.L. Meriam and L.G. Kraige, Engineering Mechanics, DYNAMICS, 5th Edition
5/1, 5/2, 5/3, 5/4 (review) , 5/5, 5/6 (review), 6/6, 6/7

12

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

J. Peraire, S. Widnall
16.07 Dynamics
Fall 2008
Version 1.0

Lecture L22 - 2D Rigid Body Dynamics: Work and Energy


In this lecture, we will revisit the principle of work and energy introduced in lecture L11-13 for particle
dynamics, and extend it to 2D rigid body dynamics.

Kinetic Energy for a 2D Rigid Body


We start by recalling the kinetic energy expression for a system of particles derived in lecture L11,

T =

1
1
2
mvG
+
mi ri 2 ,
2
2
i=1

where n is the total number of particles, mi denotes the mass of particle i, and r i is the position vector of
n
particle i with respect to the center of mass, G. Also, m = i=1 mi is the total mass of the system, and
v G is the velocity of the center of mass. The above expression states that the kinetic energy of a system of
particles equals the kinetic energy of a particle of mass m moving with the velocity of the center of mass,
plus the kinetic energy due to the motion of the particles relative to the center of mass, G.
For a 2D rigid body, the velocity of all particles relative to the center of mass is a pure rotation. Thus, we
can write
r i = r i .
Therefore, we have
n

1
i=1

mi ri 2

1
i=1

mi (

r i )

r i )

1
i=1

where we have used the fact that and r i are perpendicular. The term

mi ri 2 2 ,
n

i=1

mi ri 2 is easily recognized as

the moment of inertia, IG , about the center of mass, G. Therefore, for a 2D rigid body, the kinetic energy
is simply,
T =

1
1
2
mvG
+ IG 2 .
2
2
1

(1)

2
When the body is rotating about a xed point O, we can write IO = IG + mrG
and

T =

1
1
1
2
2
mvG
+ (IO mrG
) 2 = IO 2 ,
2
2
2

since vG = rG .

The above expression is also applicable in the more general case when there is no xed point in the motion,

provided that O is replaced by the instantaneous center of rotation. Thus, in general,

T =

1
IC 2 .
2

We shall see that, when the instantaneous center of rotation is known, the use of the above expression does
simplify the algebra considerably.
Example

Rolling Cylinder

Consider the cylinder rolling without slipping on a at plane. The friction between the plane and the
cylinder insure the no-slip condition, but the friction force does no work. The no-slip condition requires
= VG /R0 ). We take the cylinder to have its center of mass at the center of the cylinder but we allow the

mass distribution to be non-unform by allowing a radius of gyration kG . For a uniform cylinder kG = R0 / 2

The total kinetic energy, from equation (1) is given by


T =

1
1
1
1
2 2
2
mVG2 + mkG
VG /R02 = m(R02 + kG
) 2 = IC 2
2
2
2
2

(2)

where the parallel axis theorem has been used to relate the moment of inertia about the contact point C to
the moment of inertia about the center of mass.

Work
External Forces
Since the body is rigid and the internal forces act in equal and opposite directions, only the external forces
applied to the rigid body are capable of doing any work. Thus, the total work done on the body will be
n
n

(Wi )12 =
i=1

i=1

(r i )2

F i dr ,

(r i )1

where F i is the sum of all the external forces acting on particle i.


Work done by couples
If the sum of the external forces acting on the rigid body is zero, it is still possible to have non-zero work.
Consider, for instance, a moment M = F a acting on a rigid body. If the body undergoes a pure translation,
it is clear that all the points in the body experience the same displacement, and, hence, the total work done
by a couple is zero. On the other hand, if the body experiences a rotation d, then the work done by the
couple is
dW = F

a
a
d + F d = F ad = M d .
2
2

If M is constant, the work is simply W12 = M (2 1 ). In other words, the couples do work which results
in the kinetic energy of rotation.

Conservative Forces
When the forces can be derived from a potential energy function, V , we say the forces are conservative. In
such cases, we have that F = V , and the work and energy relation in equation 4 takes a particularly
simple form. Recall that a necessary, but not sucient, condition for a force to be conservative is that it
must be a function of position only, i.e. F (r) and V (r). Common examples of conservative forces are gravity
(a constant force independent of the height), gravitational attraction between two bodies (a force inversely
proportional to the squared distance between the bodies), and the force of a perfectly elastic spring.

The work done by a conservative force between position r 1 and r 2 is


r2
r
W12 =
F dr = [V ]r21 = V (r 1 ) V (r 2 ) = V1 V2 .
r1

NC
Thus, if we call W12
the work done by all the external forces which are non conservative, we can write the

general expression,
NC
T1 + V1 + W12
= T2 + V2 .

Of course, if all the forces that do work are conservative, we obtain conservation of total energy, which can
be expressed as,
T + V = constant .

Gravity Potential for a Rigid Body


In this case, the potential Vi associated with particle i is simply Vi = mi gzi , where zi is the height of particle
i above some reference height. The force acting on particle i will then be F i = Vi . The work done on
the whole body will be
n

i=1

r 2i

r 1i

F i dr i =

n
n

((Vi )1 (Vi )2 ) =
mi g((zi )1 (zi )2 = V1 V2 ,
i=1

i=1

where the gravity potential for the rigid body is simply,


V =

mi gzi = mgzG ,

i=1

where zG is the z coordinate of the center of mass. Its obvious but worth noting that because the gravitational
potential is taken about the center of mass, the inertia plays no role in determining the gravitational potential.

Example

Cylinder on a Ramp

We consider a homogeneous cylinder released from rest at the top of a ramp of angle , and use conservation
of energy to derive an expression for the velocity of the cylinder.

Conservation of energy implies that T +V = Tinitial +Vinitial . Initially, the kinetic energy is zero, Tinitial = 0.
Thus, for a later time, the kinetic energy is given by
T = Vinitial V = mgs sin ,
4

where s is the distance traveled down the ramp. The kinetic energy is simply T =

1
2
2 IC ,

where IC =

IG + mR2 is the moment of inertia about the instantaneous center of rotation C, and is the angular
velocity. Thus, IC 2 = 2mgs sin , or,
v2 =

2gs sin
,
1 + (IG /mR2 )

since = v/R. For the general case of a cylinder with the center of mass at the center of the circle but an
2
uneven mass distribution, we write T = 12 m(1 + kG
/R2 ), where the eect of mass distribution is captured

in kG ; the smaller kG , the more concentrated the mass about the center of the cylinder. Then
v2 =

2gssin
2 /R2
1 + kG

(3)

This equation shows that the more the mass is concentrated towards the center of the cylinder (kG small),
a higher velocity will be reached for a given height, i.e less of the potential energy will go into rotational
kinetic energy.

Equilibrium and Stability


If all the forces acting on the body are conservative, then the potential energy can be used very eectively
to determine the equilibrium positions of a system and the nature of the stability at these positions. Let us
assume that all the forces acting on the system can be derived from a potential energy function, V . It is
clear that if F = V = 0 for some position, this will be a point of equilibrium in the sense that if the
body is at rest (kinetic energy zero), then there will be no forces (and hence, no acceleration) to change
the equilibrium, since the resultant force F is zero. Once equilibrium has been established, the stability of
the equilibrium point can be determine by examining the shape of the potential function. If the potential
function has a minimum at the equilibrium point, then the equilibrium will be stable. This means that if
the potential energy is at a minimum, there is no potential energy left that can be traded for kinetic energy.
Analogously, if the potential energy is at a maximum, then the equilibrium point is unstable.
Example

Equilibrium and Stability

A cylinder of radius R, for which the center of gravity, G, is at a distance d from the geometric center, C,
lies on a rough plane inclined at an angle .

Since gravity is the only external force acting on the cylinder that is capable of doing any work, we can
examine the equilibrium and stability of the system by considering the potential energy function. We have
zC = zC0 R sin , where zC0 is the value of zC when = 0. Thus, since d = |CG|, we have,
V = mgzG = mg(zC + d sin ) = mg(zC0 R sin + d sin ) .
The equilibrium points are given by V = 0, but, in this case, since the position of the system is uniquely
determined by a single coordinate, e.g. , we can write
V =

dV
,
d

which implies that, for equilibrium, dV /d = mg(R sin + d cos ) = 0, or, cos = (R sin )/d. If
d < R sin , there will be no equilibrium positions. On the other hand, if d R sin , then eq. =
cos1 [(R sin )/d] is an equilibrium point. We note that if eq. is an equilibrium point, then eq. is also an
equilibrium point (i.e. cos = cos()).

In order to study the stability of the equilibrium points, we need to determine whether the potential energy
is a maximum or a minimum at these points. Since d2 V /d2 = mgd sin , we have that when eq. < 0,
then d2 V /d2 > 0 and the potential energy is a minimum at that point. Consequently, for eq. < 0, the
equilibrium is stable. On the other hand, for eq. > 0, the equilibrium point is unstable.

Example

Oscillating Cylinder and Ellipse

Consider the solid semi-circle at rest on a at plane in the presence of gravity. At rest, it is in equilibrium since
the gravitational moments balance. We consider that it tips and rolls, keeping the no-slip condition satised.
This motion results in a vertical displacement of the center of mass, y and a horizontal displacement of
the center of mass x, where y and x can be found from the geometry.

To determine the stability, we consider the change in potential energy, V (). Only the vertical displacement
of the center of mass contributes to a change in potential. If we expand the potential V () for small , we
will get an expression V () = A2 . (Recall that for the pendulum, V () = mgL2 /2.) The question of
stability depends upon the sign of A. If A is positive, the system is stable; if A is negative, the system is
unstable.
For positive A, the next step is to determine the frequency of oscillation. It is obvious that the semi-circle will
oscillate about is center of symmetry. To determine the frequency, we need to identify TM AX , the maximum
value of kinetic energy. The system has both translation and rotational kinetic energy, and both will be at
their maximum values when the system moves through the point of symmetry, = 0.
The translational kinetic energy will reect the maximum velocity of the center of mass; the rotation kinetic
energy, the maximum value of the angular velocity as the system moves through the point of symmetry.
We now consider the two systems shown in the gure. These are simply semi-ellipses resting on a at plane.

Again, the point of symmetry will be an equilibrium point since the gravitational moments will balance.
But the question of stability relates to whether the center of mass moves up or down as increases. We
have V () = A2 , with stability for A > 0 and instability for < 0. We feel instinctively, that one of these
systemsthe tall skinny one is unstable. This implies that it will not remain balanced about the equilibrium
point, but will tip over.

ADDITIONAL READING
J.L. Meriam and L.G. Kraige, Engineering Mechanics, DYNAMICS, 5th Edition
6/6, 6/7

Recall that the work done by a force, F , over an innitesimal displacement, dr, is dW = F dr. If F total
i
denotes the resultant of all forces acting on particle i, then we can write,
dWi = F total
dr i = mi
i

dv i
1
dr i = mi v i dv i = d( mi vi2 ) = d(Ti ) ,
dt
2

where we have assumed that the velocity is measured relative to an inertial reference frame, and, hence,
F total
= mi ai . The above equation states that the work done on particle i by the resultant force F total
is
i
i
equal to the change in its kinetic energy.

The total work done on particle i, when moving from position 1 to position 2, is

2
(Wi )12 =
dWi ,
1

and, summing over all particles, we obtain the principle of work and energy for systems of particles,
T1 +

(Wi )12 = T2 .

(4)

i=1

The force acting on each particle will be the sum of the internal forces caused by the other particles, and
the external forces. We now consider separately the work done by the internal and external forces.

Internal Forces
We shall assume, once again, that the internal forces due to interactions between particles act along the lines
joining the particles, thereby satisfying Newtons third law. Thus, if f ij denotes the force that particle j
exerts on particle i, we have that f ij is parallel to r i r j , and satises f ij = f ji .
Let us now look at two particles, i and j, undergoing an innitesimal rigid body motion, and consider the
term,
f ij dr i + f ji dr j .

(5)

If we write dr j = dr i + d(r j r i ), then,


f ij dr i + f ji dr j = f ij (dr i dr i ) f ij d(r j r i ) = f ij d(r j r i ).
It turns out that f ij d(r j r i ) is zero, since d(r j r i ) is perpendicular to r j r i , and hence it is also
perpendicular to f ij . The orthogonality between d(r j r i ) and r j r i follows from the fact that the distance
between any two particles in a rigid body must remain constant (i.e. (r j r i ) (r j r i ) = (rj ri )2 = const;
thus dierentiating, we have 2d(r j r i ) (r j r i ) = 0).

We conclude that, since all the work done by the internal forces can be written as a sum of terms of the
n
form 5, then the contribution of all the internal forces to the term i=1 (Wi )12 in equation 4, is zero.

10

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

J. Peraire, S. Widnall
16.07 Dynamics
Fall 2007
Version 1.0

Lecture L23 - 2D Rigid Body Dynamics: Impulse and Momentum


In lecture L9, we saw the principle of impulse and momentum applied to particle motion. This principle
was of particular importance when the applied forces were functions of time and when interactions between
particles occurred over very short times, such as with impact forces. In this lecture, we extend these principles
to two dimensional rigid body dynamics.

Impulse and Momentum Equations


Linear Momentum
In lecture L22, we introduced the equations of motion for a two dimensional rigid body. The linear momentum
for a system of particles is dened as
L = mv G ,
where m is the total mass of the system, and v G is the velocity of the center of mass measured with respect
to an inertial reference frame. Assuming the mass of the system to be constant, we have that the sum of the
= F , or, integrating
external applied forces to the system, F , must equal the change in linear momentum, L
between times t1 and t2 ,

t2

L2 L1 =

F dt

(1)

t1

Note that this is a vector equation and therefore must be satised for each component separately. Expression
1 is particularly useful when the precise time variation of the applied forces is unknown, but their total impulse
can be calculated. Of course, when the impulse of the applied forces is zero, the momentum is conserved
and we have (v G )2 = (v G )1 .

Angular Momentum
A similar expression to 1 can be derived for the angular momentum if we start from the principle of conser
G = M G . Here, H G = IG is the angular momentum about the center of
vation of angular momentum, H
mass, and M G is the moment of all externally applied forces about the center of mass. Integrating between
times t1 and t2 , we have,

t2

(H G )2 (H G )1 =

M G dt .
t1

(2)

In a similar manner, for rotation about a xed point O, we can write,


t2
(H O )2 (H O )1 =
M O dt ,

(3)

t1

where H O = IO , the moment of inertia, IO , refers to the xed point O, and the external moments are
with respect to point O.
Finally, if the external applied moment is zero, then we have conservation of angular momentum, which
implies 2 = 1 .

Example

Cylinder rolling over a step

We consider a disk of radius R rolling over a at surface, and striking a rough step of height h.

Initial Impact
First, we consider the situation an instant before and an instant after the impact with the step. Before
the impact, the velocity of the center of mass, v G , is in the horizontal direction and has a magnitude of
vG = R. After the cylinder hits the step, it starts turning around point O, (the notation O will be used
throughout this discussion for the instantaneous point of contact) and therefore the velocity of the center
of mass, v G , must be perpendicular to the segment OG. Clearly, this change in velocity direction happens
over a very short time interval and is caused by the impulsive forces generated during the impact. Since we
are assuming a rough step, the point on the cylinder which is in contact with the step at O has zero velocity
(no slipping) during this short interval. Therefore, we can use equation 3 about point O to characterize the
change in v G . The angular momentum of the cylinder about O before impact is
HO = IG + mR sin vG = [

IG
+ m(R h)] vG .
R

The angular momentum just after impact is,

HO
= IO = IO

vG
IG

=[
+ mR] vG
.
R
R

In these equations, and are the angular velocities before and after the impact, respectively, and m is
the mass of the cylinder. The change in angular momentum, according to equation 3, is equal to the angular
2

impulse about O. Clearly, the normal and tangential forces, N and F , generate no moment about O. The
only other force is the weight, W , which is not an impulsive force. Hence, the impulse generated over a very

short time interval, t2 t1 , will be negligible. Thus, HO = HO


, which implies that,

vG
=

where kO =

IG + mR(R h)
IG + mR(R h)
Rh
vG =
vG = (1 2 ) vG ,
IG + mR2
IO
kO

(4)

IO /m is the radius of gyration of the cylinder about O.

Rolling about O
After the initial impact, the cylinder rolls about point O. During this process, since there are no dissipation
mechanisms, energy will be conserved. Therefore, we can use the conservation of energy principle and require
that the sum of the potential and kinetic energies remains constant.

The change in potential energy, mgh, is obtained by a decrease in the kinetic energy, thus
1
1
2
2
IO = mgh + IO ,
2
2
or,
2

v G = v G 2gh

R2
2 .
kO

The residual velocity after the cylinder has climbed the step, vG
, can be expressed in terms of the initial

velocity, vG , using 4, as,


2

v G = (1

Rh 2 2
R2
) vG 2gh 2 .
2
kO
kO

If we want to determine the minimum initial velocity, (vG )min , that would allow the cylinder to climb the

step, we set the residual velocity, vG


, to zero and obtain,
2

(vG )2min

2gh kR2

(1

Rh 2
2 )
kO

Rebounding
So far, we have assumed that the cylinder pivots around point O without loosing contact with the step. It

is clear that if the velocity vG


is very large, then, as soon as the turning starts, a force will be required

to produce the centripetal acceleration. For small velocities, the weight should be sucient, but, for larger
velocities, it is possible that the normal reaction force, N , will become zero, and the cylinder will loose
3

contact with the step. The normal force, N , can be calculated from the normal momentum equation. That
is, the sum of the forces in the direction of OG should equal the centripetal acceleration,
2

W sin N = m

vG
.
R

Setting W = mg, sin = (R h)/R, and N = 0, we can determine the minimum velocity, (vG
)rebound , that

will produce separation. The resulting expression can be combined with equation 4 to obtain an expression
for the minimum initial velocity required for separation,
(vG )2rebound =

g(R h)
.
(1 Rh
)2
k2
O

We note that for h/(R h) <

2
kO
/(2R2 ),

(vG )min < (vG )rebound , and, in this case, there is a range of

velocities for which it is possible for the cylinder to climb the step without rebounding.

Center of Percussion Relative to an Instantaneous Center of Motion


In some situations, it is of interest to determine how a body should be impulsively set in motion such that
a certain prescribed pointthe point Cwill be (at least momentarily) the instantaneous center of motion.

Consider a rigid body of mass m which is initially at rest. An impulse J is applied at t = 0 at point P in
the body. The application of J to the body initiates both translational and rotational motion. Thus,
mv G

J ,

IG

r J .

The modulus of the initial angular velocity is = dP J/IG , and the modulus of the velocity of the center
of mass is vG = J/m. We now wish to nd out the point P such that a prescribed point C in the body
is the instantaneous center of motion. If C is the center of motion, then v G = r CG , or, in magnitude
vG = dC . Therefore,
dC =

vG
J/m
IG
=
=
,

dP J/IG
mdP

or,
dP =

IG
k2
= G .
mdC
dC
4

or
2
dC dP = kG

in other words, for any body, the product of dP the distance betweens the point of application of an impulse
and dC , the center of motion is a property of the body, the radius of gyration squared. Note that, since the
impulsive moment is a vector, the same result would be obtained if the impulse J were applied anywhere
along the line through P in the direction of J . The point P is called the center of percussion associated
with the center of motion C.

Example

Striking a billiard ball

Consider a billiard ball resting on a table. In the general case, if we apply an impulse at a point h above the
table, there will be a reaction friction impulse with the table and the ball will roll. However, this process
is unreliable, in that the coecient of friction might not be large enough to produce a smooth motion. For
improved accuracy we want to hit the ball at a height h such that no ction force is required. Therefore
we want to know at what height above the table we have to hit a billiard ball so that no friction impulse
occurs and the ball rolls on the table without slipping.

The moment of inertia of a homogeneous sphere about its center of mass is IG = (2/5)mR2 . We take
moments about the instantaneous center of rotation, the contact point C between the ball and the table. By
the parallel axis theorem, the moment of inertia about C is IC = 7/5M R02 . The governing equations are:
1) relate linear impulse to the change in linear momentum for the center of mass, M VG = P PF ; 2) relate
the moment of the impulse to the change in angular momentum taken about the point O, P h = IO ; 3) use
the geometric relationship, vG = R0 . We obtain the following result for the friction force required if the
ball is to roll smoothly on the table.
PF = P (1 h
5

5
)
7R0

(5)

therefore, if h = 75 R0 , no friction force will occur and the ball will more reliably roll smoothly. For our earlier
discussion of center of percussion, we see that the point h is the center of percussion relative to the contact
(instantaneous center of rotation) point O.

ADDITIONAL READING
J.L. Meriam and L.G. Kraige, Engineering Mechanics, DYNAMICS, 5th Edition
6/8

References
[1] M. Martinez-Sanchez, Unied Engineering Notes, Course 95-96.

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

J. Peraire, S. Widnall
16.07 Dynamics
Fall 2008
Version 1.0

Lecture L24 - Pendulums


A pendulum is a rigid body suspended from a xed point (hinge) which is oset with respect to the bodys
center of mass. If all the mass is assumed to be concentrated at a point, we obtain the idealized simple
pendulum. Pendulums have played an important role in the history of dynamics. Galileo identied the
pendulum as the rst example of synchronous motion, which led to the rst successful clock developed
by Huygens. This clock incorporated a feedback mechanism that injected energy into the oscillations (the
escapement, a mechanism used in timepieces to control movement and to provide periodic energy impulses
to a pendulum or balance) to compensate for friction loses. In addition to horology (the science of measuring
time), pendulums have important applications in gravimetry (the measurement of the specic gravity) and
inertial navigation.

Simple Pendulum
Consider a simple pendulum of mass m and length L.

The equation of motion can be derived from the conservation of angular momentum about the hinge point,
O,
IO = mgL sin .
Since the moment of inertia is simply IO = mL2 , we obtain the following non-linear equation of motion,
g
+ sin = 0 .
L

(1)

Multiplying this equation by , we can write,


d 1 2 g
( cos ) = 0,
dt 2
L
which implies that 2 (2g/L) cos = constant. Setting = max , when = 0 we have,

2g
=
(cos cos max ) .
L
This equation cannot be integrated further in an explicit manner. Its solution must be expressed in terms
of, so called, elliptic functions. The period of the oscillation, T , is obtained by multiplying by four the time
it takes for the pendulum to go from = 0 to = max . Thus,
max
4
d

T =
.
cos cos max
(2g/L) 0
Again, this is an integral which cannot be evaluated explicitly, but can be approximated, assuming that
max is not very large, as (the algebra is omitted here),

L
2
T 2
1 + max .
g
16

(2)

Small amplitude approximation


If we assume that the amplitude of pendulums oscillation is small, then sin , and the equation of
motion, given by 1, becomes linear,
g
+ = 0 .
L

(3)

This expression is much simpler than equation 1, and has solutions of the form,
= A sin n t + B cos n t ,
where n =

g/L is the natural frequency of oscillation. It is clear that these solutions are periodic, and

the period is given by


2
T =
= 2
n

L
.
g

(4)

Setting = 0 for t = 0, and = max for t = T /4, we obtain,


= max sin n t .
We observe that, in the small amplitude case, the period is independent of max . This is called synchronism
and is central to time-keeping functions in clocks. This means that, provided the amplitude is small, small
changes in amplitude due to friction or other disturbances have little eect on the period. Comparing
expression 4 with the approximate solution of the non-linear problem given by equation 2, we see that for
amplitudes of, say, max = 30 , the dierence between the two expressions is only about 1.7%.

Compound Pendulum
In the case of a compound pendulum, we can proceed in a similar manner.

Conservation of angular momentum about O gives,


IO = mgrG sin .
Expressing IO in terms of the radius of gyration, kO ,
2
IO = mkO

we have,
grG
+ 2 sin = 0 .
kO

(5)

We note that this equation is the same as equation 1 for the simple pendulum, if we identify the term g/L in
2
equation 1 with the term grG /kO
in equation 5. This leads to the denition of an equivalent length, Lequiv ,

as,
Lequiv =

2
2
kO
k 2 + rG
= G
.
rG
rG

(6)

2
2
2
2
Here we have used the fact that IO = IG + mrG
, and therefore kO
= kG
+ rG
. Thus, we have that the

motion of a compound pendulum is identical to that of a simple pendulum of equivalent length Lequiv , given
by equation 6. Using the small amplitude approximation, the period of the compound pendulum will be

Lequiv
T = 2
.
g
One question we may want to ask is whether, for a given body (kG xed), we can make the period (or Lequiv )
arbitrarily small by choosing rG (or the hinge point O) appropriately. From equation 6, we can write
Lequiv
rG
1
=
+
,
kG
kG
(rG /kG )
which shows that, when we try to reduce Lequiv /kG by reducing rG , the term rG /kG is reduced, but the
term 1/(rG /kG ) increases. This situation is shown in the graph, which also shows that the minimum value
for Lequiv /kG is 2, a value which is attained for rG = kG .

In conclusion, we have that for a given solid, the shortest equivalent length, and hence the fastest pendulum,
occurs when it is suspended from a point which is at a distance rG = kG from the center of mass. In this
case, the minimum equivalent length is
(Lequiv )min = 2kG .

Example

Minimum period pendulum

Consider a uniform rod of length l. In this case, IG = ml2 /12, and kG = l/ 12 0.29l. Therefore, the
fastest pendulum is obtained when the bar is suspended from a distance 0.29l away from the center of mass.
We note that if the suspension point is moved slightly, the period of the pendulum will increase. However,
moving the suspension point will make practically no dierence in the frequency, because the tangent at point
A (see graph) is horizontal. This fact has been used in the construction of extremely accurate pendulums
for clocks.

Center of Percussion
The center of percussion associated with a given center of motion was introduced in lecture L24. In the case
of a pendulum, the desired center of motion is the hinge, O. The center of percussion is on the extension of
the line that connects O with the center of mass.

From lecture L24, we know that the distance, rP , between the center of percussion, P , and the center of
mass, G, is given by,
rP =

2
kG
.
rG

(7)

Recall that the center of percussion is the point at which we should strike the pendulum with a horizontal
impulse so that O becomes the instantaneous center of motion. It is clear that, if O is the instantaneous
center of motion, the horizontal reaction needed to keep the pendulum in place will be zero and, as a
consequence, the impulse will not be felt at O. In fact, this can be veried directly in a straightforward
manner. Let us assume that the pendulum is given an impulse J at a distance rP below the center of mass,
G. Conservation of momentum in the horizontal direction and conservation of angular momentum around
G imply that,
mvG

= J JR

IG

= JrP + JR rG .

Solving for JR , we have that,


JR = J

2
IG mrG rP

kG
rG rP
=
J
,
2
IG + mr
G
IO

which clearly shows that the impulse reaction, JR , is zero when rP satises equation 7.

We note that, for a given body, rP will be large when rG is very small, and, as a consequence, the center of

percussion will not be a material point in the body.

Reversibility
It turns out that if the pendulum is suspended from point P , the roles of rG and rP are reversed. In
particular, point O is the center of percussion relative to the instantaneous center of motion, P . Using

equations 6 and 7, the equivalent length of a compound pendulum can be written as,
Lequiv =

2
2
2
kG
+ rG
rP rG + rG
=
= rG + rP ,
rG
rG

and the corresponding period is,

T = 2

Lequiv
= 2
g

rG + rP
.
g

It should be clear that when the pendulum is suspended from point P , the equivalent length does not change,
since,
2
kG
+ rP2
rP rG + rP2
=
= rG + rP Lequiv .
rP
rP

The Kater Pendulum


The reversibility of point O and P described above is the principle of the reversible pendulum, invented by
Kater to measure gravity with high accuracy. Katers pendulum consists of a long bar, equipped with two
xed knife edges at an accurately known distance L, and with some moveable masses positioned along the
bar. The positions of these masses on the bar are adjusted until the periods associated with suspension from
either knife edge are precisely equal to, say, T . This guarantees that each point is the center of percussion
relative to the other, and, thus, L = Lequiv . The local gravity is then given by,
2
2
g = Lequiv
.
T
The Schuler Pendulum
Consider a pendulum suspended in vertical position from point O. If the support is suddenly accelerated in
the horizontal direction, the pendulum will rotate due to the inertial forces acting on its center of mass. It
is not dicult to show that the instantaneous center of motion in this case will be precisely the center of
percussion associated with O.
Now, imagine a pendulum on the earths surface whose center of percussion is at the earths center. When
the pendulums support is accelerated, the pendulum rotates in a way such that it always points towards the
center of the earth. Such a pendulum always keeps itself vertical regardless of the acceleration, and would
be of obvious interest for applications such as inertial navigation.
It is interesting to see that, although conceptually correct, the construction of such a pendulum would pose
serious technological challenges. If we assume an earth radius of R = 6370 km, then rP R. Considering a
pendulum with, say, kG = 0.3m, the distance between the support and the center of mass would have to be,
rG =

0.32
= 1.5 108 m = 0.015m.
6.37 106

This distance approaches atomic dimensions and, in fact, such a pendulum has never been constructed
using purely mechanical means. It is, however, routinely realized by replacing the physical pendulum with
a combination of gyroscope and accelerometers having the same dynamics.
6

It is interesting to note that the period of the Schuler pendulum is given by

Lequiv
R
TSchuler = 2
2
84.4min. ,
g
g
which is exactly the same period as that of a circular orbit around the earth at zero altitude.

References
[1] M. Martinez-Sanchez, Unied Engineering Notes, Course 95-96.
[2] J. P. Den Hartog, Mechanics, Dover, 1961

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

J. Peraire, S. Widnall
16.07 Dynamics
Fall 2008
Version 2.0

Lecture L25 - 3D Rigid Body Kinematics


In this lecture, we consider the motion of a 3D rigid body. We shall see that in the general three-dimensional
case, the angular velocity of the body can change in magnitude as well as in direction, and, as a consequence,
the motion is considerably more complicated than that in two dimensions.

Rotation About a Fixed Point


We consider rst the simplied situation in which the 3D body moves in such a way that there is always a
point, O, which is xed. It is clear that, in this case, the path of any point in the rigid body which is at a
distance r from O will be on a sphere of radius r that is centered at O. We point out that the xed point
O is not necessarily a point in rigid body (the second example in this notes illustrates this point).
Eulers theorem states that the general displacement of a rigid body, with one xed point is a rotation about
some axis. This means that any two rotations of arbitrary magnitude about dierent axes can always be
combined into a single rotation about some axis.
At rst sight, it seems that we should be able to express a rotation as a vector which has a direction along
the axis of rotation and a magnitude that is equal to the angle of rotation. Unfortunately, if we consider two
such rotation vectors, 1 and 2 , not only would the combined rotation be dierent from 1 + 2 , but in
general 1 + 2 =
2 + 1 . This situation is illustrated in the gure below, in which we consider a 3D rigid
body undergoing two 90o rotations about the x and y axis.

It is clear that the result of applying the rotation in x rst and then in y is dierent from the result obtained
by rotating rst in y and then in x. Therefore, it is clear that nite rotations cannot be treated as vectors,
since they do not satisfy simple vector operations such as the parallelogram vector addition law.
This result can also be understood by considering the rotation of axes by a coordinate transformation.
Consider a transformation (x ) = [T1 ](x), and a subsequent coordinate transformation (x ) = [T2 ](x ). The
(x ) coordinate obtained by the transformation (x ) = [T1 ][T2 ](x) will not be the same as the coordinates
obtained by the transformation (x ) = [T2 ][T1 ](x), in other words, order matters.

Angular Velocity About a Fixed Point


On the other hand, if we consider innitesimal rotations only, it is not dicult to verify that they do
indeed behave as vectors. This is illustrated in the gure below, which considers the eect of two combined
innitesimal rotations, d 1 and d 2 , on point A.

Image by MIT OpenCourseWare.

(gure reproduced from J.L. Meriam and K.L. Kraige, Dynamics, 5th edition, Wiley)

As a result of d 1 , point A has a displacement d 1 r, and, as a result of d 2 , point A has a displacement


d 2 r. The total displacement of point A can then be obtained as d r, where d = d 1 + d 2 . Therefore,
it follows that angular velocities 1 = 1 and 2 = 2 can be added vectorially to give = 1 + 2 . This
means that if at any instant the body is rotating about a given axis with angular velocity 1 and at the
same time this axis is rotating about another axis with angular velocity 2 , the total angular velocity of
the body will be simply = 1 + 2 . Therefore, although the nite rotations of a body about an axis are

not vectors, the innitesimal rotations are vectors. The angular velocity is thus a vector and for a complex
conguration, the various components can ba vectorially added to obtain the total angular velocity.
Consider the complex rotating conguration shown below. We want to determine the angular velocity of the
disc D.

First, we note that the disc is rotating with angular velocity 1 about the axis M M . In turn, this axis is
rotating with angular velocity 2 about the horizontal axis, which is at this instant aligned with the x axis.
At the same time, the whole assembly is rotating about the z axis with angular velocity 3 . Therefore,
the total angular velocity of the disc is vector sum of the individual angular velocity vectors. The resultant
vector is shown in the gure: total = 1 + 2 + 3 . Expressed in the xed x, y, z system for which the
conguration is instantaneously aligned as shown, we have
= 2 i + 1 cos j + (1 sin + 3 ) k .
Here, is the angle between M M and the y axis.

Angular Acceleration
In order to apply the principle of conservation of angular momentum, we need a general expression for
the angular acceleration of the various components of a complex rotating conguration. We rst examine
this problem with respect to inertial coordinate system, x, y, z. In order to apply conservation of angular
=
momentum, we must develop a formula for the time rate of change of angular momentum, BH
d
IG dt

d
dt
IG .

We will later consider under what conditions we need to consider

For now, we concentrate on determining

d
dt .

d
dt (IG )

d
dt
IG .

We rst consider a common situation in the study of rotating

3D objects, the rotating wheel with angular velocity attached to a central hub which rotates with angular
velocity , shown in the gure. The total angular velocity is the vector sum of and .
3

In this example, we take constant in magnitude (but not direction) and as constant. It is clear that
the rotation will rotate the vector , changing its direction. The magnitude of is , the direction
is normal to ; by Coriolis theorem, the result is = . It is interesting to note that this result is
independent of the distance b between the wheel and the axis of rotation for . This is a consequence of our
earlier observation that in a rigid body rotating with angular velocity , every point rotates with angular
velocity .
Example

Multiple Observers

In this example we illustrate a more systematic procedure for calculating the angular velocities and accel
erations when several reference frames are involved. We want to determine the angular acceleration of the
disc D as a function of the angular velocities and accelerations given in the diagram. The angle of 1 with
the horizontal is .

The angular velocity vector for each component is the vector sum of the individual angular velocities of
the components: 3 , 2total = 3 + 2 ; 1total = 1 + 2 + 3 . The angular accelerations of the various
components are worked through individually. Consider 3 . Since it is the primary rotor, the rotation 2
and 1 do not aect its motion. Therefore the angular acceleration is due solely to 3 . Consider 2 . This
angular velocity vector will change with time both due to 2 as well as 3 2 . And nally, for 1 : this
angular velocity vector will change with time both due to 1 as well as ( 3 + 2 ) 1 .
We now resolve these vectors into appropriate coordinate systems. We consider three sets of axes. Axes
xyz are xed. Axes x y z rotate with angular velocity 3 k with respect to xyz. Axes x y z rotate with
angular velocity 2 i with respect to x y z . Finally, the disc rotates with angular velocity 1 j with respect
to the axes x y z .
The angular velocity of the disc with respect to the xed axes will be simply
= 2 i + 1 j + 3 k .

(1)

At the instant considered, i = i and j = cos j + sin k. Therefore, we can also write,
= 2 i + 1 cos j + (1 sin + 3 )k .
In order to calculate the angular acceleration of the disc with respect to the xed axes xyz, we start from
(1) and write,

d
dt

=
xyz

d
d
(2 i + 1 j + 3 k)
=
(2 i + 1 j )
+ 3 k .
dt
dt
xyz
xyz

Here, we have used the fact that k does not change with respect to the xyz axes and therefore only the
magnitude of 3 changes. In order to calculate the time derivative of 2 i + 1 j with respect to the inertial
reference frame, we apply Coriolis theorem. Since x y z rotates with angular velocity 3 k with respect to
5

xyz, we write,

d
d

(2 i + 1 j )
=
(2 i + 1 j )
+ 3 k (2 i + 1 j ) .
dt
dt
xyz
x y z

In the x y z frame, i does not change direction. Therefore, we can write

d
d
(2 i + 1 j )
= 2 i +
(1 j )
+ 3 k (2 i + 1 j ) .
dt
dt

xyz
xy z
In order to evaluate the derivative of 1 j with respect to the x y z frame we make use again of Coriolis
theorem, and write

d
d

(1 j )
=
(1 j )
+ 2 i 1 j = 1 j + 2 i 1 j .
dt
dt
x y z
x y z
Here, we have used the fact that x y z rotates with angular velocity 2 i with respect to x y z , and the
derivative of j in the x y z reference frame is zero. Putting it all together, we have,

d
= 2 i + 1 j + 2 i 1 j + 3 k (2 i + 1 j ) + 3 k
dt xyz
= 2 i + 1 j + 1 2 k + 2 3 j 1 3 cos i + 3 k

( 2 1 3 cos ) i + ( 1 cos + 2 3 1 2 sin ) j + ( 3 + 1 sin + 1 2 cos ) k.

Instantaneous Axis of Rotation


In two dimensions, we introduced the concept of instantaneous center of rotation. For a rotating body with
one xed point, we can extend this concept to an instantaneous axis of rotation.
Consider a vertical disc rotating with a constant angular velocity , rolling without slip with an angular
velocity about the vertical axis with angular velocity . The no-slip condition requires that the velocity of
the center of mass VG = R. The disk rolling around a circle of radius b will have an angular velocity
= R/b. We note that just as in two-dimensions, point O is a xed point, an instantaneous center of
rotation. That is, at that instant, the motion of the disc is such that the distance from any point in the disc
to the point O remains constant. The total instantaneous angular velocity is the vector sum of + . The
resultant vector is sketched and as would be expected, it is parallel to a line from the center of the hub to
the point C.
Once the instantaneous angular velocity, T = + , has been determined, the velocity of any point in the
rigid body is simply
v = T r ,

(2)

where r is the position vector of the point considered with respect to the xed point O. It follows that for
any point which is on the line passing through O and parallel to T , the velocity will be zero. This line is
therefore the Instantaneous Axis of Rotation.
6

We can now dene two space curves, which are instantaneously tangent to the instantaneous axis of rotation.
These curves are useful for more complex problems in the general rotation of a body about a xed point.
This simple example gives us a useful visualization of these space curves. The rst is called the body cone:
it is the locus of points made by the instantaneous axis of rotation as the body traces its motion. As can be
seen, for this case, it is a cone of radius equal to the radius of the disc, extending to the hub at the z axis.
(The term body cone is a bit of a misnomer; the body doesnt have to t entirely inside the body cone.)
The second space curve is called the space cone. It is dened as the locus of points traced in space by the
instantaneous axis of rotation. At the instantaneous position of the body, these two curves are both tangent
to the instantaneous axis of rotation. In this case, the space cone is the cone bounded by the curve traced
by the instantaneous contact point and the xed point, the center of the hub. The body curve and the space
curve for more general three-dimensional motions are shown in the gure.

For a general motion, as the direction of the instantaneous axis of rotation (or the line passing through O
parallel to ) changes in space, the locus of points dened by the axis generates a xed Space Cone. If the
change in this axis is viewed with respect to the rotating body, the locus of the axis generates a Body Cone.
7

At any given instant, these two cones are tangent along the instantaneous axis of rotation. When the body

is in motion, the body cone appears to roll either on the inside or the outside of the xed space cone. This

situation is illustrated in the gure below.

The acceleration of any point in the rigid body is obtained by taking the derivative of expression 2. Thus,

a = T r + T r = r + T ( T r) .

(3)

Here, is the angular acceleration vector and is locally tangent to both the Space and the Body Cones.

General Motion
In the general case, the displacement of a rigid body is determined by a translation plus a rotation about

some axis. This result is a generalization of Eulers theorem, which is sometimes known as Chasles theorem.

In practice, this means that six parameters are needed to dene the position of a 3D rigid body. For instance,

we could choose three coordinates to specify the position of the center of mass, two angles to dene the axis

of rotation and an additional angle to determine the magnitude of the rotation.

Unlike the motion about a xed point, it is not always possible to dene an instantaneous axis of rotation.

Consider, for instance, a body which is rotating with angular velocity and, at the same time, has a

translational velocity parallel to . It is clear that, in this case, all the points in the body have a non-zero

velocity, and therefore an instantaneous center of rotation cannot be dened.

It turns out that, in some situations, the motion of the center of mass of a 3D rigid body can be determined

independent of the orientation. Consider, for instance, the motion of an orbiting satellite in free ight. In

this situation, the sum of all external forces on the satellite does not depend on the satellites attitude, and,

therefore, it is possible to determine the position without knowing the attitude. In more complex situations,

however, it may be necessary to solve simultaneously for both the position of the center of mass and the

attitude.

The velocity, v P , and acceleration, aP , of a point, P , in the rigid body can be determined if we know the

velocity, v O , and acceleration, aO , of a point in the rigid body, O , as well as the bodys angular velocity,

, and acceleration, .

vP

= v O + r P

(4)

aP

= aO + r P + ( r P ) .

(5)

Here, r P is the position vector of the point, P , relative to O . We point out that the angular velocity and
angular acceleration are the same for all the points in the rigid body.

ADDITIONAL READING
8

J.L. Meriam and L.G. Kraige, Engineering Mechanics, DYNAMICS, 5th Edition
7/1, 7/2, 7/3, 7/4, 7/5, 7/6 (review)

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

J. Peraire, S. Widnall
16.07 Dynamics
Fall 2008
Version 2.1

Lecture L26 - 3D Rigid Body Dynamics: The Inertia Tensor


In this lecture, we will derive an expression for the angular momentum of a 3D rigid body. We shall see that
this introduces the concept of the Inertia Tensor.

Angular Momentum
We start from the expression of the angular momentum of a system of particles about the center of mass,
H G , derived in lecture L11,
HG =

(r i mi ( r i )) =

i=1

HG =

(r i

mi ri 2

(1)

i=1

mi (

r i ))

i=1

mi ri 2

i=1

r v dm

(2)

r v dm .

HG =
m

Here, r is the position vector relative to the center of mass, v is the velocity relative to the center of mass.
We note that, in the above expression, an integral is used instead of a summation, since we are now dealing
with a continuum distribution of mass.

For a 3D rigid body, the distance between any particle and the center of mass will remain constant, and the
particle velocity, relative to the center of mass, will be given by
v = r .
1

Thus, we have,

r ( r ) dm =

HG =

[(r r ) (r )r ] dm .
m

Here, we have used the vector identity A (B C) = (A C)B (A B)C. We note that, for planar bodies
undergoing a 2D motion in its own plane, r is perpendicular to , and the term (r ) is zero. In this case,

the vectors and H G are always parallel. In the three-dimensional case however, this simplication does

not occur, and as a consequence, the angular velocity vector, , and the angular momentum vector, H G ,

are in general, not parallel.

In cartesian coordinates, we have, r = x i + y j + z k and = x i + y j + z k, and the above expression

can be expanded to yield,

HG

(x + y + z ) dm

(x x + y y + z z )x dm i

2
2
2


+
y
(x + y + z ) dm
(x x + y y + z z )y dm j
m
m

2
2
2


+
z
(x + y + z ) dm
(x x + y y + z z )z dm k
m

( Ixx x Ixy y Ixz z ) i

(Iyx x + Iyy y Iyz z ) j

(Izx x Izy y + Izz z ) k .

(3)

The quantities Ixx , Iyy , and Izz are called moments of inertia with respect to the x, y and z axis, respectively,
and are given by

Ixx =

(y 2 + z 2 ) dm ,

Iyy =

(x2 + z 2 ) dm ,

Izz =

(x2 + y 2 ) dm .

We observe that the quantity in the integrand is precisely the square of the distance to the x, y and z axis,
respectively. They are analogous to the moment of inertia used in the two dimensional case. It is also clear,
from their expressions, that the moments of inertia are always positive. The quantities Ixy , Ixz , Iyx , Iyz , Izx
and Izy are called products of inertia. They can be positive, negative, or zero, and are given by,

Ixy = Iyx =
x y dm ,
Ixz = Izx =
x z dm ,
Iyz = Izy =
y z dm .
m

They are a measure of the imbalance in the mass distribution. If we are interested in calculating the angular
momentum with respect to a xed point O then, the resulting expression would be,
HO

= ( (Ixx )O x (Ixy )O y (Ixz )O z ) i


+

((Iyx )O x + (Iyy )O y (Iyz )O z ) j

((Izx )O x (Izy )O y + (Izz )O z ) k .

(4)

Here, the moments of products of inertia have expressions which are analogous to those given above but
with x , y and z replaced by x, y and z. Thus, we have that

(Ixx )O =
(y 2 + z 2 ) dm ,
(Iyy )O =
(x2 + z 2 ) dm ,
m

(x2 + y 2 ) dm ,

(Izz )O =

and,

(Ixy )O = (Iyx )O =

xy dm ,

(Ixz )O = (Izx )O =

xz dm ,

(Iyz )O = (Izy )O =

yz dm .
m

The Tensor of Inertia


The expression for angular momentum given by equation (3), can be written in matrix form as,

HGx

HGy

HGz

Ixx



= Iyx

Izx

Ixy

Ixz

Iyy

Iyz

Izy

Izz

y .

(5)

or,
H G = [IG ],

(6)

where [IG ] is the tensor of inertia (written in matrix form) about the center of mass G and with respect to
the xyz axes. The tensor of inertia gives us an idea about how the mass is distributed in a rigid body.
Analogously, we can dene the tensor of inertia about point O, by writing equation(4) in matrix form. Thus,
we have
H O = [IO ] ,
where the components of [IO ] are the moments and products of inertia about point O given above.
It follows from the denition of the products of inertia, that the tensors of inertia are always symmetric. The
implications of equation (5) are that in many situations of importance, even for bodes of some symmetry, the
and the angular velocity vector
are not parallel. This introduces considerable
angular momentum vector H
complexity into the analysis of the dynamics of rotating bodies in three dimensions.

Principal Axes of Inertia


For a general three-dimensional body, it is always possible to nd 3 mutually orthogonal axis (an x, y, z
coordinate system) for which the products of inertia are zero, and the inertia matrix takes a diagonal form.
In most problems, this would be the preferred system in which to formulate a problem. For a rotation
about only one of these axis, the angular momentum vector is parallel to the angular velocity vector. For
symmetric bodies, it may be obvious which axis are principle axis. However, for an irregular-shaped body
this coordinate system may be dicult to determine by inspection; we will present a general method to
determine these axes in the next section.
But, if the body has symmetries with respect to some of the axis, then some of the products of inertia
become zero and we can identify the principal axes. For instance, if the body is symmetric with respect to
the plane x = 0 then, we will have Ix y = Iy x = Ix z = Iz x = 0 and x will be a principal axis. This can
be shown by looking at the denition of the products of inertia.

The integral for, say, Ix y can be decomposed into two integrals for the two halves of the body at either side
of the plane x = 0. The integrand on one half, x y , will be equal in magnitude and opposite in sign to the
integrand on the other half (because x will change sign). Therefore, the integrals over the two halves will
cancel each other and the product of inertia Ix y will be zero. (As will the product of inertia Ix z )
Also, if the body is symmetric with respect to two planes passing through the center of mass which are
orthogonal to the coordinate axis, then the tensor of inertia is diagonal, with Ix y = Ix z = Iy z = 0.

Another case of practical importance is when we consider axisymmetric bodies of revolution. In this case,
if one of the axis coincides with the axis of symmetry, the tensor of inertia has a simple diagonal form. For
an axisymmetric body, the moments of inertia about the two axis in the plane will be equal. Therefore,
the moment about any axis in this plane is equal to one of these. And therefore, any axis in the plane is
a principal axis. One can extend this to show that if the moment of inertia is equal about two axis in the
plane (IP P = Ixx ), whether or not they are orthogonal, then all axes in the plane are principal axes and the
moment of inertia is the same about all of them. In its inertial properties, the body behaves like a circular
cylinder.

The tensor of inertia will take dierent forms when expressed in dierent axes. When the axes are such that
the tensor of inertia is diagonal, then these axes are called the principal axes of inertia.

The Search for Principal Axes and Moments of Inertia as an Eigenvalue Problem
Three orthogonal principal axes of inertia always exist even though in bodies without symmetries their
directions may not be obvious. To nd the principle axis of a general body consider the body shown in the
gure that rotates about an unknown principal axis. The the total angular momentum vector is I
in the
direction of the principle axis. For rotation about the principal axis, the angular momentum and the angular
velocity are in the same direction.

We seek a coordinate axes x, y and z, about which a rotation x , y and z , which is aligned with this
coordinate direction, will be parallel to the angular momentum



HGx
Ix
I 0



HGy = Iy = 0 I



HGz
Iz
0 0

vector and related by the equation

0
x

0 y .

I
z

(7)

We then express the general form for angular momentum vector in components along the x, y and z axis

in term of the components of


along these axes using the general form of the inertia tensor in the x, y, z

system, we have

HGx

HGy

HGz

Ixx



= Iyx

Izx

To obtain the special directions of that



HGx
Ix
I



HGy = Iy = 0



HGz
Iz
0

Ixy

Ixz

Iyz y .

Izz
z

Iyy
Izy

is aligned with

0 0

I 0 y

0 I
z

(8)

a principal axis, we equate these two


Ixx
Ixy Ixz

= Iyx Iyy
Iyz y

Izx Izy Izz


z

expressions.

(9)

At this point in the process we know the inertia tensor in an arbitrary x, y, and z system and are seeking the
special orientation of which will align the angular momentum HG with the angular velocity . Collecting
terms from equation(11) on the left-hand side, we obtain

(I I) Ixy
Ixz
xx

Iyx
(Iyy I) Iyz

Izx
Izy
(Izz I)
resulting in the requirement that

I
Ixy
xx

Iyx Iyy

Izx Izy

Ixz

Iyz I 0

Izz
0
6

0
1
0


y = 0 .


z
0

(10)





0 y = 0


1
z
0

(11)

The structure of the solution for nding the principal axes of inertia and their magnitudes is a characteristicvalue problem. The three eigenvalues give the directions of the three principal axis, and the three eigen
vectors give the moments of inertia with respect to each of these axis.
In principal directions, the inertia tensor has the form

I
0
x

[IG ] = 0 Iy

0 0

Iz

where we will write Ix = Ixx , Iy = Iyy and Iz = Izz . Also, in principal axes we will then have
H G = Ix x i + Iy y j + Iz z k .

Parallel Axis Theorem


It will often be easier to obtain the tensor of inertia with respect to axis passing through the center of mass.
In some problems however, we will need to calculate the tensor of inertia about dierent axes. The parallel
axis theorem introduced in lecture L22 for the two dimensional moments of inertia can be extended and
applied to each of the components of the tensor of inertia.

In particular we can write,

(Ixx )O

((yG + y )2 + (zG + z )2 ) dm

2
2

2
2
=
(y + z ) + 2yG
y dm + 2zG
z dm + (yG + zG )
dm
=

(y 2 + z 2 ) dm =

2
2
= Ixx + m(yG
+ zG
).

Here, we have use the fact that y and z are the coordinates relative to the center of mass and therefore
their integrals over the body are equal to zero. Similarly, we can write,
2
(Iyy )O = Iyy + m(x2G + zG
),

2
(Izz )O = Izz + m(x2G + yG
),

and,
(Ixy )O = (Iyx )O = Ixy + mxG yG ,

(Ixz )O = (Izx )O = Ixz + mxG zG ,


7

(Iyz )O = (Izy )O = Iyz + myG zG .

Rotation of Axes
In some situations, we will know the tensor of inertia with respect to some axes xyz and, we will be interested
in calculating the tensor of inertia with respect to another set of axis x y z . We denote by i, j and k the
unit vectors along the direction of xyz axes, and by i , j and k the unit vectors along the direction of x y z
axes. The transformation of the inertia tensor can be accomplished by considering the transformation of the
and the angular velocity vector
and
angular momentum vector H
. We begin with the expression of H

in the x1 , x2 , x3 system.
= [I]
H

(12)

We have not indicated by a subscript where the origin of our coordinates are, i.e. the center of mass G,
a xed point O, or any other point, because as long as we are simply doing a rotational transformation of
coordinates about this point, it does not matter,
From Lecture 3, we have that the transformation of a vector from a coordinate system x1 , x2 , x3 into a
coordinate system x1 , x2 x3

H
1

H2

H3

is given by

i i
1 1

= i2 i1

i3 i1

i1 i2

i1 i3

i2 i2

i2 i3 H2 = [T ] H2 .

i3 i3
H3
H3

i3 i2

H1

where we have introduced the symbol [T] for the transformation matrix.

H1

Likewise, the transformation of


into the

i i
1 1 1

2 = i2 i1


3
i3 i1

x1 , x2 , x3 system is given by

i1 i2 i1 i3
1
1

i2 i2 i2 i3 2 = [T ] 2 .

i3 i2 i3 i3
3
3

(in the transformed system) we multiply both sides of equation (10) by [T ], and obtain
To determine H
= [T ]H
= [T ][I]
H

(13)

where [I] is the inertia matrix in the original coordinate system. But this equation is not quite in the proper
= [I ]
form; we need to have the form H
to identify [I ], the inertia matrix in the new coordinate system.
We take advantage of the fact that the matrix [T ]T is also the inverse of [T ], [T ]1 , so that [T ]T [T ] = [Iden],
where [Iden] is the identity matrix. Now, we can always stick an identity matrix in a matrix equationlike
multiplying by 1so that we may write

= [T ]H
= [T ][I][T ]T [T ]
H
= [T ][I][T ]T [T ]
= [I ]

(14)

were we have grouped the terms so that it is obvious that the inertia tensor in the transformed coordinate
system is
[I ] = [T ][I][T ]T

(15)

Therefore, if we want to calculate the tensor of inertia with respect to axis x y z , we can write in matrix
form

I
xx

Iy x

Iz x

Ix y
I

y y

Iz y

Ix z
I

y z

Iz z

i i



= j i

k i

i j

j j

k j

i k

Ixx

j k Iyx

k k
Izx

Ixy

Ixz

Iyy

Iyz

Izy

Izz

i i

j i

i j

i k

j k

cos (12 )

cos (13 )

ji

k i

kj .

k k

where we have returned to the x, y, z notation.

Likewise, expressed in the direction cosines of the gure, the transformation is

cos (11 ) cos (12 ) cos (13 )


Ixx Ixy Ixz
cos (11 )

[I ] = cos (21 ) cos (22 ) cos (23 ) Iyx Iyy Iyz cos (21 )

cos (31 ) cos (32 ) cos (33 )


Izx Izy
Izz
cos (31 )

cos (22 )
cos (32 )

cos (23 ) ..

cos (33 )
(16)

Inertia Tensor of a Cube


We are going to consider and calculate the inertia tensor of a cube of equal length sides b. (This example is
taken from Marion and Thorton.) For this choice of coordinates, it is obvious that the center of mass does
not lie at the origin. It is also obvious that there is considerable symmetry in the geometry.

Evaluating the various integrals for the components of the inertia tensor, we have

2
1
1
2
2
2
M
b

M
b

M
b
4
4
3

1
2
1
2
2
2
[I] = 4 M b

M
b

M
b
3
4

1
1
2
2
2
2
4Mb 4Mb
M
b
3

(17)

This result is perhaps a bit surprising; because of the strong symmetry of the cube we might have expected
these coordinates to be principal axes. However, our origin of coordinates is not at the center on mass, and
by the parallel axis theorem, we incur products of inertia the equal to total mass times the various distances
of center of mass from the origin of coordinates.
There are several options to nd a set of principal axis for this problem. One is to use the symmetry to
identify a coordinate system in which the products of inertia vanish. The transformation sketched in the
gure is a good candidate to obtain the principal axes; that is a rotation of = /4 about the x3 axis,
followed by a rotation of = sin1 13 about the x1 axis. This would result in the x1 axis going through the
upper point of the cube and through the center of mass as well. If we were to do this, we would nd that
indeed this transformation results in coordinate directions which are principal axis directions, and moreover
that the inertias are equal about any axis in the plane perpendicular to the diagonal of the cube, so that all
these axes are principal with equal moments of inertia.
Another more general method is to express the search for principal axes as an eigenvalue problem, as previ
ously outlined. In a problem without strong and obvious symmetry, identifying an appropriate transformation
is often a matter of guess work; therefore the eigenvalue formulation is more useful.

10

Products of Inertia in a Plane


In many situations, we will know one principal axis from symmetry, and will be working with moments of
inertia and inertia products in a plane. An example would be a at plate of mass m, thickness t, width a and
length b. It is clear from the sketch that the principal axis are the x y coordinates. We may have reasons
for wanting to express the moments of inertia and the products of inertia in the x , y system (rotated from
the x, y system by an angle ) or to move back and forth between the x, y system, and the x , y system.

The coordinate transformation that takes a vector in


x
cos



y = sin


z
0

the x, y, z system

sin 0
x

cos 0 y

0
1
z

into the x , y , z system is

(18)

or (x ) = [T ](x); while the coordinate transformation that takes a vector in the x , y , z system into the
x, y, z system is

sin

cos


y = sin


z
0

cos
0

0 y .

1
z

(19)

or (x) = [T ]T where we have taken advantage of the fact that the matrix for the reverse transformation
(x to x) is the transpose of the matrix for the forward transformation (x to x ). The inertia matrix for a
rectangular plate with its center of mass at the origin in the x, y, z system, for which the principal axes line
up with the coordinate directions is

b2
12

Ix,y,z = M 0

a2
12

a2 +b2
12

(20)

If we wish to obtain the inertia tensor in the x , y , z system, we apply the transformation
Ix ,y ,z = [T ][Ix,y,z ][T ]T
11

(21)

The Symmetry of Three


An interesting and practical result for inertias in a plane comes from considering the moment of inertia of a
body with tri-symmetry, such as a three-bladed propeller or a wind turbine. It is clear that the axis normal
to the gure is a principal axis, with masses symmetrically balanced about the origin. It is also clear that
both the x and y axis are principal axis, with inertia products going to zero by symmetry. What is perhaps
surprising, is that Ixx the moment of inertia about the x axis is equal to Iyy the moment of inertia about
the y axis.

Consider the x axis; for a point r0 the contribution to the moment of inertia from all 3 mass points is
Ixx = m(r02 + 2(r0 /2)2 ) = m

3r02
2 .

Consider the y axis; for a point r0 the contribution to the moment of

3r 2
inertia from all 3 mass points is Iyy = m(2( 3r0 /2)2 ) = m 20 . Therefore, the total moment of inertia about
both the x and y axis will be an integral of the mass distribution in dm times

3r02
2 .

Since the moments of

inertia are equal about the 2 principal axis x and y, the moment of inertia about any axis in the x,y plane
is also equal to Ixx . The three-bladed propeller behaves as a uniform disk. This has important implications
for the vibrations of propellers of historically-interesting high performance propeller ghter aircraft and wind
turbines in yawing motion.
ADDITIONAL READING
J.L. Meriam and L.G. Kraige, Engineering Mechanics, DYNAMICS, 5th Edition
7/7, Appendix B
S. T. Thorton and J. B. Marion, Classical Dynamics of Particles and Systems, 4th Edition, Chapter 11

12

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

J. Peraire, S. Widnall
16.07 Dynamics
Fall 2008
Version 2.0

Lecture L27 - 3D Rigid Body Dynamics: Kinetic Energy;

Instability; Equations of Motion

3D Rigid Body Dynamics


In Lecture 25 and 26, we laid the foundation for our study of the three-dimensional dynamics of rigid bodies
by: 1.) developing the framework for the description of changes in angular velocity due to a general motion
of a three-dimensional rotating body; and 2.) developing the framework for the eects of the distribution of
mass of a three-dimensional rotating body on its motion, dening the principal axes of a body, the inertia
tensor, and how to change from one reference coordinate system to another.
We now undertake the description of angular momentum, moments and motion of a general three-dimensional
rotating body. We approach this very dicult general problem from two points of view.
The rst is to prescribe the motion in term of given rotations about xed axes and solve for the force system
required to sustain this motion.
The second is to study the free motions of a body in a simple force eld such as gravitational force acting
through the center of mass or free motion such as occurs in a zero-g environment. The typical problems
in this second category involve gyroscopes and spinning tops. The second set of problems is by far the more
dicult.
Before we begin this general approach, we examine a case where kinetic energy can give us considerable
insight into the behavior of a rotating body. This example has considerable practical importance and its
neglect has been the cause of several system failures.

Kinetic Energy for Systems of Particles


In Lecture 11, we derived the expression for the kinetic energy of a system of particles. Here, we derive the
expression for the kinetic energy of a system of particles that will be used in the following lectures. A typical
particle, i, will have a mass mi , an absolute velocity v i , and a kinetic energy Ti = (1/2)mi v i v i = (1/2)mi vi2 .
The total kinetic energy of the system, T , is simply the sum of the kinetic energies for each particle,
T =

i=1

Ti =

1
i=1

mi vi2 =

1
1
2
mvG
+
mi vi 2 , .
2
2
i=1

where v is the velocity relative to the center of mass. Thus, we see that the kinetic energy of a system of
particles equals the kinetic energy of a particle of mass m moving with the velocity of the center of mass,
plus the kinetic energy due to the motion of the particles relative to the center of mass, G.
We have said nothing about the conservation of energy for a system of particles. As we shall see, that
depends upon the details of internal interactions and the work done by the external forces. The same is
true for a rigid body. Work done by internal stresses, or energy lost due to the complexities of a system
described as a rigid body, but which in reality may have internal modes which drain energy, will act to
decrease the kinetic energy. Thus, although we can condently relate angular momentum to external forces,
we have no such condence in dealing with conservation of energy. As we shall see, this has important
technological implications, especially for the stability of spacecraft. In this section we consider the particles
to be components of a rigid body.

Kinetic Energy for a 3D Rigid Body


For a rigid body, the summation i = 1, n becomes an integral over the total mass M.

1 2
1
1 2
2
T =
v dm = M vG
+
v dm .
2
m 2
m 2
For a rigid body, the velocity relative to the center of mass is written
v =
r

(1)

where r is the vector to the mass dm for the center of mass G.

Using the vector identity


B
) C
=A
(B
C
),
(A

(2)

we have
v 2 = v v = (
r ) (
(r (
r )) =
r v )
.
Therefore,

v 2 dm =

G,
r v dm =
H

(3)

and the total expression for kinetic energy of a rigid body can be written
T =

1
2
G = 1 vG .L
G + 1
G
M vG
+
H
H
2
2
2

(4)

G is the linear momentum.


where M L
If there is a xed point O about which the body rotates, then
T =

1
0

H
2

0 is the angular momentum about point O. Even if there is no xed point about which the body
where H
rotates, there is an instantaneous center of rotation C. In some situations, it may be useful to write
TC =

1
C

H
2

(5)

If the angular velocity is expressed in principal axes, then the angular momentum about the center of mass
G = Ixx xi + Iyy yj + Izz zk so that the kinetic energy can be written
can be written H
T =

1
1
2
M vG
+ (Ixx x2 + Iyy y2 + Izz z2 )
2
2

(6)

The above expressions for kinetic energy are useful to apply the principle of work and energy. We see that
2
the kinetic energy has two components: one is due to the translation, 12 M vG
. The other is due to rotation.

The rotational component can be written as

TR =

1
G = 1 (Ixx x2 + Iyy y2 + Izz z2 )

H
2
2

(7)

We have no assurance that energy will be conserved. Internal motions can dissipation energy. This is not true
for angular momentum: internal forces and motions will not change angular momentum. The implications
of this can be profound.
Consider a body spinning about the z axis of inertia with no external moments. Then the angular momentum
G = Hzk will be
will be constant and since it is a vector, it will always point in the z direction so that H
constant. Now let us assume that Izz < Ixx and Izz < Iyy . In this case, for a given angular momentum,
the kinetic energy is a maximum, since T = 12 Hz2 /Izz , and Izz is the smallest moment of inertia. If there is
any dissipation mechanismstructural damping, fuel sloshing, frictionthen the angular momentum will stay
constant in magnitude and direction while the energy decreases. The only way, that the energy can decrease
while keeping the angular momentum constant is to change the axis of spin to one with a larger moment
of inertia while keeping the vector direction and magnitude of the angular momentum vector constant.
Therefore, if the body tumbles and begins to spin about the x or y body-xed axis, the kinetic energy
becomes T = 12 Hz2 /(Ixx orIyy ). The kinetic energy will be reduced since both Ixx and Iyy are greater than
Izz .

The sketch show the process by which this would occur. An initial spin about the bodys z axis will transition
G constant in magnitude and direction.
over time to a spin about the bodys x or y axis, while keeping the H
The initial state is simple; the nal state is simple; the intermediate states are complex in that the body
axis is not aligned with the rotation vector, much like a spinning top performing a complex motion. Do we
have to worry about this?
The rst American satellite, Explorer 1, was built by the Jet Propulsion Laboratory in Pasadena, and
launched from Canaveral on 31 January 1958 by a modied Jupiter-C missile into an orbit ranging between
354 kilometers and 2,515 kilometers at an angle of 33 degrees to the equator. The satellite was long and
slender. It was spin-stabilized about its long axisthe one with the smallest moment of inertia. Spin
stabilization has advantages in that the angular momentum remains xed in direction in space under passive
control, if its stable! The satellite had exible antenna which vibrated and dissipated energy. During the
mission, it unexpectedly suered an attitude failure in that it tipped over and began to spin about the axis
with the maximum moment of inertia. Other space systems as well as re-entry vehicles are prone to this
phenomena.

Governing Equations for Rotational Motion of a Three-Dimensional


Body
The governing equations are those of conservation of linear momentum L = M v G and angular momentum,
H = [I], where we have written the moment of inertia in matrix form to remind us that in general the
direction of the angular momentum is not in the direction of the rotation vector . Conservation of linear
momentum requires
=F
L

(8)

Conservation of angular momentum, about a xed point O, requires


0=M
H

(9)

G = MG
H

(10)

or about the center of mass G

We rst examine cases in which the motion is specied and our task is to determine the forces and moments.

Example: Rotating Skew Rod


We consider a simple example which illustrate the eect of rotation about a non-principal axis. In this case,
we consider two masses equal m1 and m2 attached to a massless rigid rod of length 2l aligned from the
horizontal by an angle and rotating about the z axis with constant angular velocity k. We examine this
problem at the instant when the rod is aligned with the x and z axis. Because of symmetry, the center of
mass is at the origin. Therefore, we consider only the change in angular momentum. Because the magnitude
of the angular velocity is constant, we are concerned only with its change in direction.

The angular momentum for this system is H =

r mi v i . For this case, the velocity for mass m1 is

v 1 = lcosj, while for mass m2 the velocity is v 2 = lcosj while r 1 = lcosi + lsink and r 2 =
lcosi lsink. At this instant of time, the angular momentum is in the x, z plane and of magnitude
H = 2ml2 cossini + 2ml2 cos2 k

(11)

Since the axis about which the rotation takes place is not a principal axis, it is no surprise that the angular
momentum vector is not aligned with the angular velocity vector. In this simple case, since we are dealing
with two mass points, the rod has zero moment of inertia when rotate about is axis. Therefore the angular
momentum vector is perpendicular to the rod. As the skew rod rotates about the z axis, the angular
momentum changes as
= H = 2m 2 l2 sincosj
H

(12)

resulting in a required moment about the y axis to sustain the motion of


M = 2m 2 l2 sincosj

(13)

This moment has a physical interpretation in term of centripetal acceleration. As the rod rotates with
angular velocity , the individual mass points experience a centripetal acceleration directed towards the
z axis of a = 2 lcosi. This requires an external force directed along the i direction: on m1 , a force of
F 1 = 2 mlcosi; on m2 , a force of F 2 = 2 mlcosi ; no net force is required because the body is supported
at its center of mass. These forces can only be supplied/resisted by an external moment applied at the origin,
where the rod is supported, a moment about the origin of M = r 1 F 1 + r 2 F 2 = 2m 2 l2 sincosj,
which is the moment required to sustain the motion.
As the rod rotates about the z axis, the moment or torque required to sustain the motion also rotates. Thus
any xture that is used to attach the rotating rod to the axle must sustain this torque.
Observe that this force goes to zero for = 0 and = /2. For these s, the rod would be rotating about
a principal axis.

Example: Rotating Cylinder


The previous discussion can easily be extended, replacing the rod by a cylinder. In this case the moment of
inertia Ix x is not equal to 0. Therefore, the angular momentum vector is not perpendicular to the cylinder
but is at an angle determined by the relative values of Iz z and Ix x . If the body is not a cylinder, i.e.
Ix x =
Iy y =
Iz z , then the result is qualitatively similar but the angular momentum vector H is not in
the plane formed by the cylinder axis and .

Example: Use of the inertia tensor


These examples may be treated using the more formal description in term of the inertia tensor. The principal
axis of this system is clearly along the rod joining the two masses, as shown in the gure. If we treat the
masses as mass points, there is no moment of inertia about the x axis, while the moments of inertia about
the y and z axis are equal and equal to Iz z = Iy y = 2ml2 . The coordinate transformation from the
principal axes x , y to the x, y axis is


x
cos


y = 0


z
sin

0 sin
1


y .

0 cos

(14)

The inertia tensor in the x, y, z system is obtained by the transformation introduced in Lecture 26: [I] =
[T ][I ][T ]T .

I
Ixy
xx

Ixy Iyy

Ixz Iyz

Ixz

cos



Iyz = 0

Izz
sin

0 sin
1

0 cos

0 2ml2

0 0

resulting in [I] for the inertia matrix in the x, y, z system

sin2

[I] = 2ml2 0

cossin

cos

2
2ml
sin

0 sin
1

0 cos

(15)

as
0

cossin

0 cos2

(16)

This conrms that the x and z

HGx

HGy = 2ml2

HGz

axis are not principal axes. The angular momentum vector is then

sin2
0 cossin
0
cossin

0 = 2ml2

0
1 0
0

2
2
cossin 0 cos

cos

(17)

for the angular momentum. Leading to the required moment to sustain the motion as
= H = 2l2 m 2 cos sin j
M =H

(18)

in agreement with equation (13).

Example: Two-Bladed Wind Turbines


An exciting application of the analysis of the rotating skew rod is to the operation of a two-bladed wind
turbine. Recall that the analysis predicted that a constant moment about a horizontal axis is required from
the support structure if the skew rod is to rotate about the vertical axis with angular velocity . This
analysis is also valid if is a function of time if we add the angular momentum about the y axis as a result of
the propeller rotating about its own axis. Let = t, that is the rod rotates about the horizontal y axis
through its center of mass and perpendicular to the plane containing the rod and masses, like a propeller.
In addition, it rotates about the z axis following the previous analysis of the rotating skew rod. The total
angular momentum is then the result from the previous analysis with = t; the angular momentum of
the skew rod becomes H s = 2ml2 cos(t)sin(t)i + 2ml2 cos2 (t)k plus angular momentum due to the
rotation of the propeller of magnitude H p = 2ml2 j. The additional change in angular momentum due to
p = H p . This is constant in time in a coordinate system rotating
the propeller rotation is simply H
with the plane of the propeller. The motion requires an applied steady moment.

The total angular momentum of this conguration is then the sum of that for the skew rod plus that for
the rotation of the propeller about the y axis. However, the additional change of angular momentum occurs
only through the action of k in rotating the angular momentum of the propeller. The angular velocity of
the propeller, does not rotate the angular momentum vector of the skew rod.
H = H s + H p = 2l2 mcos(t)sin(t)i 2ml2 j + 2ml2 cos2 (t)k

(19)

Since only k rotates the angular momentum vector, we have from Coriolis theorem that the rate of change
of angular momentum is
= (k) (H s + H p )
H

(20)

M = 2l2 m(i 2 cos(t)sin(t)j + 0k.

(21)

Then the moment required is

Note the unsteady behavior of the required moment with time. The moment required to produce the
propeller rotation about j, and to move the direction of this rotation about the z axis with angular velocity
is steady; the moment required to produce the rotation of the skew rod about k is unsteady. This is due
to the imbalance in inertia. At various times in its motion, the device is rotating about the z axis ( k) which
is not a principal axis for the conguration.
Consider this problem as a model for a two-bladed wind turbine. The wind turbine will happily rotate about
its y axis with angular velocity . But wind turbines must yaw to face into a new wind direction. This
requires for a short time an angular velocity about the z axis. From our analysis, as the wind turbine
9

rotates in yaw to face the new wind direction, an oscillatory moment of M = 2l2 m 2 cos(t) sin(t)j will
be exerted on the support. This undesirable oscillatory load will be eliminated if the wind turbine has three
blades so that all directions in the x, z plane are principal axis. The gure shows how this idea has caught
on.

Image by brentdanley on Flickr.

ADDITIONAL READING
J.L. Meriam and L.G. Kraige, Engineering Mechanics, DYNAMICS, 5th Edition
5/1, 5/2, 5/3, 5/4 (review) , 5/5, 5/6 (review)

10

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

J. Peraire, S. Widnall
16.07 Dynamics
Fall 2009
Version 3.0

Lecture L28 - 3D Rigid Body Dynamics: Equations of Motion;

Eulers Equations

3D Rigid Body Dynamics: Eulers Equations


We now turn to the task of deriving the general equations of motion for a three-dimensional rigid body.
These equations are referred to as Eulers equations. The governing equations are those of conservation of
linear momentum L = M v G and angular momentum, H = [I], where we have written the moment of
inertia in matrix form to remind us that in general the direction of the angular momentum is not in the
direction of the rotation vector . Conservation of linear momentum requires
=F
L

(1)

Conservation of angular momentum, about a xed point O, requires


0=M
H

(2)

G = MG
H

(3)

or about the center of mass G

In our previous application of these equations, we specied the motion and used the equations to specify
what moments would be required to produce the prescribed motion. In this more general formulation, we
allow the body to execute free motions, possibly under the action of external moments. We consider the
general motion of a body about its center of mass, rst examining this in a inertial reference frame.

At an instant of time, we can calculate the angular momentum of the body as H = [I]. One possible
method to obtain the moments and the motion of the body is to perform our analysis in this inertial
coordinate system. We would of course align our coordinate system initially with the principal axes of the
body. We could then write
G = d/dt([I]) = [I] + [I]
MG = H

(4)

This would be a appropriate approach but the diculty is keeping track of [I] in the inertial coordinate
system. The initial inertial axis, even if principal axis, will not remain principal axis, and the inertia seen
in this coordinate system will vary with time. So unless we are considering the motion of a sphere, for
which all axis are principal and the inertia tensor is constant about all axis, we cannot get very far with this
approach.

Body-Fixed Axis
We formulate the governing equations of motion in an axis system xed to the body, paying the price for
keeping track of the motion of the body in order to have the inertia tensor remain independent of time in
our reference frame. Given our earlier discussion of terms added to the description of motion in a rotating
and accelerating coordinate system, it may seem surprising that we can do this easily, but the statement of
conservation of angular moment about the center of mass, or about a xed point of instantaneous rotation
holds if we include the changes in angular momentum arising from Coriolis Theorem. The general body is
shown in the gure. We x the x, y, z axis to the body and instantaneously align them with x, y, z. Referring

to the gure, we see the components of ,- 1 , 2 and 3 - and the components of the angular moment
vector H, which in general is not aligned with the angular velocity vector. We also see the vectors 2 j + 3 k
applied to the x axis, with corresponding components of sketched at the y and z axes. At this instant
plus the eect of the instantaneous
of time, the change in H will be due to actual time rate of change H
rotation of the x axis due to : the change in H due to the instantaneous rotation of the coordinate system
is
= [I] + H
H

(5)

Equating the change in angular momentum to the external moments, we have the statement of conservation
of angular momentum in body-xed axes.

Mx

H x Hy z + Hz y

(6)

My

H y Hz x + Hx z

(7)

Mz

H z Hx y + Hy x

(8)

This particular form of the equations of motion is valid for any set of body-xed axes. If the axes chosen are
principal axes, then we may express the conservation of angular momentum in terms of moments of inertia
about the principle axes,

Mx

Ixx x (Iyy Izz )y z

(9)

My

Iyy y (Izz Ixx )z x

(10)

Mz

Izz z (Ixx Iyy )x y

(11)

These equations are called Eulers equations. They provide several serious challenges to obtaining the general
solution for the motion of a three-dimensional rigid body. First, they are non-linear (containing products of
the unknown s). This means that elementary solutions cannot be combined to provide the solution for a
more complex problem. But a more fundamental diculty, is that we do not know the location of the
axis system, x, y and z. Recall that since the axes are xed to the body, we are committed to follow the
body as it rotates in order to use these equations to obtain a solution. Thus, we must develop a method
to follow the changes in axis location as the body rotates. Before turning to this problem, we examine a
situation where we know the location of the axis, at least approximately.

Stability of Free Motion about a Principal Axis


Consider a body rotating about the z axis-a principal axis- with angular velocity z . Without loss of
generality, we may consider this to be a rectangular block. We take advantage of the fact that we know to
3

a good approximation the axis of rotation, at least initially. We examine the question of stability to small
perturbations, rotations about the x and y axis of magnitude x and y where x << z and y << z .
The moments of inertia about the x, y and z axes are Ixx , Iyy and Izz ; we say nothing about the magnitudes
of these inertias at this point.

Assume a small impulsive moment that initiates a small rotation about the x and y axes and thereafter the
motion proceeds with no applied external moments. For this case, Eulers equations become
0

Ixx x (Iyy Izz )y z

(12)

Iyy y (Izz Ixx )z x

(13)

Izz z (Ixx Iyy )x y

(14)

The z equation contains the product of two small terms in contrast to the other two equations where the size
of the terms is comparableas far as we can tell. We therefore note that since Izz z = (Ixx Iyy )x y with
both x and y being small quantities, we may take z as constant, equal to some . We now dierentiate
both equations with time, and substitute to obtain an equation for x . (We could do this as well for y with
the same result.)
Ixx
x

(Iyy (Izz )(Izz Ixx ) 2


x = 0
Iyy

(15)

or

x Ax = 0

(16)

The solution to this dierential equation for x (t) is

x (t) = Be

At

+ Ce

At

(17)

The stability of the motion is determined by the sign of A. If A is positive, an exponential divergence
will result, and the initial small perturbation in x will grow without bound, as least as predicted by small
4

perturbation analysis. If the sign of A is negative, oscillatory motion of x (and y ) will result, and the
motion is stable.
Examining A =

(Iyy Izz )(Izz Ixx )


,
Iyy

we see that the condition for A to be positive is that Izz is intermediate

between Ixx and Iyy . That is Ixx < Izz < Iyy or Iyy < Izz < Ixx . We conclude that a body rotating about
an axis where the moment of inertia is intermediate between the other two inertias, is unstable. Also, that
rotation about either the largest or smallest inertias is stable. This consideration relates to stability of a
rotating body as predicted from Eulers equation; we have already examined the stability of rotation about
the smallest inertia axis and concluded that if energy is dissipated, that motion is unstable.

Stability of a Gyrostat
The analysis of the proceeding section can easily be extended to a more important, more useful and more
complex conguration: the gyrostat satellite. A gyrostat consists of a spinning body which contains within
itself another spinning body, referred to as the rotor. Gyrostat satellite are used when the external body of
the satellite must spin slowly to accomplish its mission while it needs the stabilization provided by faster
rotation. This is accomplish by placing a rotor inside the satellite. The angular momentum of the rotor is
driven by a motor attached to the satellite; no net momentum increase occurs when an adjustment in rotor
speed is made. The examination of this device is quite straightforward if the rotor principal axes are aligned
with the satellite principal axes, the rotor is axisymmetric, and the center of mass of the rotor is placed at
the center of mass of the satellite. Thus the rotor is constrained to move with the rotational motion of the
satellite and principal axes remain principal.

The moment of inertia of the rotor about its principal

IR
xx

[I R ] = 0

axes is
0

R
Ixx

R
Izz

while the moment of inertia of the satellite platform is

IP 0
xx

P
[I P ] = 0
Ixx

0
0

0
0
P
Izz

(18)

(19)

Because of the geometric constraints on the system, x and y are equal for both platform and rotor while
z dier for platform, zP and rotor, zR . Since we are in a rotating coordinate system with respect to the
platform so the Ixx and Iyy are constant, zR is the relative rotation velocity between platform and rotor.
Therefore the total angular momentum is

P
Ixx

[H] = 0

P
Iyy

P
Izz

zP

P
Ixx



+ 0

0

P
Iyy

P
Izz

zR

Because of the constraints on , we can write the angular momentum as

P
R
Ixx
+ Ixx
0
0
P

P
P
R
[H] = 0
y
Iyy
+ Ixx
0

P
R R
0
0
Izz
+ Izz
z /zP
zP

(20)

(21)

where has been identied as P .


This is a remarkable result. It indicates that the eect of the rotation on the angular momentum of the
gyrostat can be incorporated by a modication of the moment of inertia about the z axis. We will call this
the eective moment of inertia,
P
R R
Izzef f = Izz
+ Izz
z /zP

(22)

where, zR is the relative rotation velocity between platform and rotor.


With this result, and the identication of the eective moment of inertia in z, the previous analysis of the
stability of rotation of a body with unequal moments of inertia goes through with the replacement of Izz by
Izzef f .
ADDITIONAL READING
J.L. Meriam and L.G. Kraige, Engineering Mechanics, DYNAMICS, 5th Edition 7.9
J. B. Marion, S. T. Thornton, Classical Dynamics of Particles and Systems, 11.8
W.T. Thompson, Introduction to Space Dynamics, Chapter 5
F. P. J. Rimrott, Introductory Attitude Dynamics, Chapter 11
6

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

J. Peraire, S. Widnall
16.07 Dynamics
Fall 2009
Version 2.0

Lecture L29 - 3D Rigid Body Dynamics

3D Rigid Body Dynamics: Euler Angles


The diculty of describing the positions of the body-xed axis of a rotating body is approached through
the use of Euler angles: spin , nutation and precession shown below in Figure 1. In this case we
surmount the diculty of keeping track of the principal axes xed to the body by making their orientation
the unknowns in our equations of motion; then the angular velocities and angular accelerations which
appear in Eulers equations are expressed in terms of these fundamental unknowns, the positions of the
principal axes expressed as angular deviations from some initial positions.
Euler angles are particularly useful to describe the motion of a body that rotates about a xed point, such
as a gyroscope or a top or a body that rotates about its center of mass, such as an aircraft or spacecraft.
Unfortunately, there is no standard formulation nor standard notation for Euler angles. We choose to follow
one typically used in physics textbooks. However, for aircraft and spacecraft motion a slightly dierent one
is used; the primary dierence is in the denition of the pitch angle. For aircraft motion, we usually refer
the motion to a horizontal rather than to a vertical axis. In a description of aircraft motion, would be the
roll angle; the yaw angle; and the pitch angle. The pitch angle would be measured from the
horizontal rather than from the vertical, as is customary and useful to describe a spinning top.

Figure 1: Euler Angles

In order to describe the angular orientation and angular velocity of a rotating body, we need three angles.

As shown on the gure, we need to specify the rotation of the body about its spin or z body-xed axis,

the angle as shown. This axis can also precess through an angle and nutate through an angle .

To develop the description of this motion, we use a series of transformations of coordinates, as we did in

Lecture 3. The nal result is shown below. This is the coordinate system used for the description of motion

of a general three-dimensional rigid body described in body-xed axis.

To identify the new positions of the principal axes as a result of angular displacement through the three

Euler angles, we go through a series of coordinate rotations, as introduced in Lecture 3.

We rst rotate from an initial X, Y, Z system into an x , y , z system through a rotation about the Z, z
axis. The angle is called the angle of precession.

x
cos sin 0
X
X

y = sin cos 0 Y = [T1 ] Y .

z
0
0
1
Z
Z
The resulting x , y coordinates remain in the X, Y plane. Then, we rotate about the x axis into the
x , y , z system through an angle . The x axis remains coincident with the x axis. The axis of rotation
for this transformation is called the line of nodes. The plane containing the x , y coordinate is now tipped
through an angle relative to the original X, Y plane. The angle is called the angle of nutation.

x
1
0
0
x
x

y = 0 cos sin y = [T2 ] y .

z
0 sin cos
z
z
And nally, we rotate about the z , z system through an angle into the x, y, z system. The z axis is called
the spin axis. It is coincident with the z axis. The angle is called the spin angle; the angular velocity
the spin velocity.

cos


y = sin


z
0

sin
cos
0

0 y

1
z

The nal Euler transformation is


x
X
coscos cossinsin


y = [T3 ][T2 ][T1 ] Y = sincos cossincos


z
Z
sinsin

= [T3 ] y

cossin + coscossin
sinsin + coscoscos
cossin

sinsin

shown. The individual coordinate rotations , and give us the angular velocities. However, these vectors
do not form an orthogonal set: is along the original Z axis; is along the line of nodes or the x axis; while

sincos Y .

cos
Z

This is the nal x, y, z body-xed coordinate system for the analysis, with angular velocities x , y , z as

is along the z or spin axis.

This is easily reorganized by taking the components of these angular velocities about the nal x, y, z coor
dinate system using the Euler angles, giving
x = sin sin + cos

(1)

y = sin cos sin

(2)

z = cos +

(3)

We could press on, developing formulae for angular momentum, and changes in angular momentum in this
coordinate system, applying these expressions to Eulers equations and develop the complete set of governing
dierential equations. In general, these equations are very dicult to solve. We will gain more understanding
by selecting a few simpler problems that are characteristic of the more general motions of rotating bodies.

3D Rigid Body Dynamics: Free Motions of a Rotating Body


We consider a rotating body in the absence of applied/external moments. There could be an overall gravi
tational force acting through the center of mass, but that will not aect our ability to study the rotational
motion about the center of mass independent of such a force and the resulting acceleration of the center of
mass. (Recall that we may equate moments to the rate of change of angular momentum about the center of
mass even if the center of mass is accelerating.) Such a body could be a satellite in rotational motion in orbit.
The rotational motion about its center of mass as described by the Euler equations will be independent of
its orbital motion as dened by Keplers laws. For this example, we consider that the body is symmetric
such that the moments of inertia about two axis are equal, Ixx = Iyy = I0 , and the moment of inertia about
z is I. The general form of Eulers equations for a free body (no applied moments) is

Ixx x (Iyy Izz )y z

(4)

Iyy y (Izz Ixx )z x

(5)

Izz z (Ixx Iyy )x y

(6)

For the special case of a symmetric body for which Ixx = Iyy = I0 and Izz = I these equations become
0

I0 x (I0 I)y z

(7)

I0 y (I I0 )z x

(8)

= I z

(9)

We conclude that for a symmetric body, z , the angular velocity about the spin axis, is constant. Inserting

this result into the two remaining equations gives


I0 x = ((I0 I)z )y

(10)

I0 y = ((I0 I)z )x .

(11)

Since z is constant, this gives two linear equations for the unknown x and y . Assuming a solution of the
form x = Ax eit and y = Ay eit , whereas before we intend to take the real part of the assumed solution,
we obtain the following solution for x and y
x = A cos t

(12)

y = A sin t

(13)

where = z (I I0 )/I0 and A is determined by initial conditions. Since z is constant, the total angular

velocity = x2 + y2 + z2 = A2 + z2 is constant. The example demonstrates the direct use of the Euler
equations. Although the components of the vector can be found from the solution of a linear equation,
additional work must be done to nd the actual position of the body. The body motion predicted by this
solution is sketched below.

The x, y, z axis are body xed axis, rotating with the body; the solutions for x (t), y (t) and z give the
components of following these moving axis. If angular velocity transducers were mounted on the body to
measure the components of , x (t), y (t) and z from the solution to the Euler equations would be obtained,
shown in the gure as functions of time. Clearly, as seen from a xed observer this body undergoes a complex
spinning and tumbling motion. We could work out the details of body motion as seen by a xed observer.
(See Marion and Thornton for details.) However, this is most easily accomplished by reformulating the
problem expressing Eulers equation using Euler angles.

Description of Free Motions of a Rotating Body Using Euler Angles


The motion of a free body, no matter how complex, proceeds with an angular momentum vector which is
constant in direction and magnitude. For body-xed principle axis, the angular momentum vector is given
by H G = Ixx x + Iyy y + Izz z . It is convenient to align the constant angular momentum vector with
the Z axis of the Euler angle system introduced previously and express the angular momentum in the i, j, k
system. The angular momentum in the x, y, z system, H G = {Hx , Hy , HZ } is obtained by applying the
Euler transformation to the angular momentum vector expressed in the X, Y, Z system, H G = {0, 0, HG }.

H G = HG sin sin i + HG sin cos j + HG cos k

(14)

Then the relationship between the angular velocity components and the Euler angles and their time derivative
given in Eq.(1-3) is used to express the angular momentum vector in the Euler angle coordinate system.

HG sin sin

= Ixx x = Ixx (sin


sin + cos )

(15)

HG sin cos

= Iyy y =

Iyy (sin
cos sin )

(16)

HG cos

= Izz z =

Izz ( cos + )

(17)

where HG is the magnitude of the H G vector.


The rst two equation can be added and subtracted to give expressions for and . Then a nal form for
spin rate can be found, resulting in
cos2 sin2
= HG (
+
)
Ixx
Iyy
1
1
= HG (

) sin sin cos


Ixx
Iyy

= HG (

1
cos2 sin2

) cos
Izz
Ixx
Iyy

(18)
(19)
(20)

For constant HG , these equations constitute a rst order set of non-linear equations for the Euler angle ,
and . In the general case, these equations must be solved numerically.
and and their time derivatives ,
Considerable simplication and insight can be gained for axisymmetric bodies for which Ixx = Iyy = I0 and
Izz = I. In this case, we have

= HG /I0

(21)

= 0

(22)

1
1
= HG ( ) cos
I
I0

(23)

Thus, in this case, the nutation angle is constant; the spin velocity is constant, and the precession
velocity is constant. Note that if I0 is greater that I, and are of the same sign; and if I0 is less than I,
and are of opposite signs; for I0 = I, the problem falls apart since we are now dealing with the inertial
equivalent of a sphere which will not exhibit precession.
We now examine the geometry of the solution in detail. We assume that the body has some initial angular
momentum that could have arisen from an earlier impulsive moment applied to the body or from a set of
initial conditions set by an earlier motion. In this case, we would know both the magnitude of the angular
momentum HG and its angle from the spin axis. The geometry of the solution is shown below. We align
the angular momentum vector with the Z axis.

For general motion of an axisymmetric body, the angular momentum H G and the angular velocity vectors
are not parallel. Using the spin axis z as a reference, the angular momentum H G makes an angle with the
spin axis; the angular velocity makes an angle with the spin axis. For this body, although the angular
momentum is not aligned with the angular velocity, they must be in the x, z plane; this is a consequence
of symmetry: that Iyy = Ixx = I0 . Therefore both y and HGy must be zero. We examine the system in
the plane formed by H g and , as shown in the gure. The relation between the angular velocities and
8

must be such that the vectors I0 y j + Iz k = H G . Since H G makes an angle with the z axis, and
makes an angle with the Z axis, we have
tan = HGx /HGz = (I0 x )/(Iz ) = (I0 /I)tan.

(24)

What we would see in this motion is a body spinning about the z axis with and precessing about the
Z axis with this is essentially a denition of precession. The vector, which is the instantaneous axis of
rotation, is in the x, z plane.
We now focus on the Z axis, and investigate the motion that can exist for a freely rotating body with the
given parameters when it has an angular momentum in the Z direction. We see that a body rotating about
its own axis with angular velocity at an angle from the Z axis, will also undergo a steady precession of
about the Z axis, keeping the angular momentum H G directed along the Z axis. The instantaneous axis
of rotation maintains a xed angle from the spin axis z. The vector maintains a xed angle from
the vertical. The motion is that of the body cone rolling around the space cone.
A variety of useful and general relations can be written between the various components of angular velocity
and angular momentum H, the Euler angles and and the angles and . Using the notation HG and

for the magnitude of the H G and vectors, we have


x =

sin

z =

cos

HG =
Hx2 + Hz2 = (x I0 )2 + (z I)2 = I02 sin2 + I 2 cos2

(sin2 + (I/I0 )2 cos2 )


=
=

(1 (I/I0 )) cos

(25)
(26)
(27)
(28)
(29)
(30)

Two dierent solution geometries are shown: one for which I0 > I; one for which I0 < I. Consider rst the
case where I < IO . This would be true for a long thin body such as a spinning football, a reentering slender
missile or an F-16 in roll.
In this case, the spin and the precession are of the same sign. The body precesses about the angular
momentum vector while spinning. The vector is the instantaneous axis of rotation. The instantaneous
axis of rotation is the instantaneous tangent between the body cone and the space cone. The outside of
the body cone, shown in the gure as a cone of half-angle aligned along the body axis, rotates about the
outside of the space cone, shown in the gure as a cone of half-angle with its axis aligned along the
angular momentum vector. This is called direct precession. For direct precession, < .
Now consider the case where I > I0 ; this would be true for a frisbee, a at spinning satellite, or a silver
dollar tossed into the air.

10

In this case, the spin and the precession are of opposite signs. This means that the vector is on the
other side of the angular momentum vector relative to the spin axis z. The body still precesses about the
angular momentum vector while spinning. The vector is still the instantaneous axis of rotation. But now,
the inside of the body cone rotates about the outside of the space cone. The angle is greater than the
angle . The body cone shown in the gure as a cone of half-angle aligned along the body axis; the space
cone has a half-angle and is aligned along the angular momentum vector. This is called retrograde
procession; the rotations are in opposite directions. Depending upon body geometry, one of these solutions
would be obtained whenever the angular momentum vector is not directed along a single principle axis. This
would occur if an impulsive moment is applied to a body along any axis which is not a principle axis. An
example of this is discussed below.

Extreme Aircraft Dynamics


The dynamics of aircraft have traditionally been dominated by aerodynamic forces. The location of the
center of mass relative to the aerodynamic center was an important consideration as was the question of
whether the vertical tail provided enough yaw moment to keep the vehicle ying straight. The details of the
response of aerodynamic forces to small disturbances of the vehicle in pitch, roll and yaw, determined the
stability of the aircraft and the frequency of the various longitudinal and lateral stability modes.
With the introduction of high-performance ghter aircraft, whose moments of inertia about all three axes
were comparable, which had the ability to initiate rapid rolling, and whose roll axis was not a principal axis,
we entered into a new ight regime where, at the limit, dynamics dominated aerodynamics.
Consider this limit, where dynamics dominates aerodynamics. We have a F-16 at high altitude where
atmospheric density is small, and we have a test pilot giving a strong roll input about the aircraft roll axis
which is not a principal axis. For simplicity, we take the y and z inertias to be equal and equal to I0 . Low
aspect ratio aircraft have moments of inertia in roll that are less than that in pitch or yaw so that this
example is given by the case I0 > I and we may apply the free-body spinning solution just discussed. We
consider that an impulsive moment/torque is applied about the roll axis and inquire what free-body motion
this would set up. At this limit, we are neglecting aerodynamics forces. In agreement with our previous
analysis, the moment about the x axis would produce an angular momentum about the x axis. But since
the x axis is not a principal axis, we would initiate a coning motion as shown.

11

This could come as quite surprise to a pilot. The question of what happens next is dependent upon the
details of the aerodynamic forces, but just to comment on the historical record: in the rst days of testing
high-performance ghter aircraft, several test pilots lost control of their aircraft, in some cases with fatal
results. This phenomenon, called roll coupling, is now well understand and incorporated into the design and
testing of new ghter aircraft.
ADDITIONAL READING
J.L. Meriam and L.G. Kraige, Engineering Mechanics, DYNAMICS, 5th Edition 7/9

W.T. Thompson, Introduction to Space Dynamics, Chapter 5

J. H. Ginsberg, Advanced Engineering Dynamics, Second Edition, Chapter 8

J.B Marion and S.T. Thornton, Classical Dynamics, Chapter 10

J.C. Slater and N.H. Frank, Mechanics, Chapter 6

12

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

J. Peraire, S. Widnall
16.07 Dynamics
Fall 2008
Version 2.0

Lecture L30 - 3D Rigid Body Dynamics: Tops and Gyroscopes

3D Rigid Body Dynamics: Euler Equations in Euler Angles


In lecture 29, we introduced the Euler angles as a framework for formulating and solving the equations for
conservation of angular momentum. We applied this framework to the free-body motion of a symmetrical
body whose angular momentum vector was not aligned with a principal axis. The angular moment was
however constant. We now apply Euler angles and Eulers equations to a slightly more general case, a top
or gyroscope in the presence of gravity.
We consider a top rotating about a xed point O on a at plane in the presence of gravity. Unlike our
previous example of free-body motion, the angular momentum vector is not aligned with the Z axis, but
precesses about the Z axis due to the applied moment. Whether we take the origin at the center of mass G
or the xed point O, the applied moment about the x axis is Mx = M gzG sin, where zG is the distance to
the center of mass..

Initially, we shall not assume steady motion, but will develop Eulers equations in the Euler angle variables
(spin), (precession) and (nutation).

Referring to the gure showing the Euler angles, and referring to our study of free-body motion, we have the
following relationships between the angular velocities along the x, y, z axes and the time rate of change of
the Euler angles. The angular velocity vectors for , and are shown in the gure. Note that these three
angular velocity vectors are not orthogonal, giving rise to some cross products when the angular velocities
i are calculated about the three principal axes.
x = sin sin + cos

(1)

y = sin
cos sin

(2)

z =

cos
+

(3)

3D Rigid Body Dynamics: Eulers Equations


We consider a symmetric body, appropriate for a top, for which the moments of inertia Ixx = Iyy = I0 and
Izz = I. The angular momentum is then
Hx = I0 x

(4)

Hy = I0 y

(5)

Hz = Iz

(6)

For the general motion of a three-dimensional body, we have Eulers equations in body-xed axes which
rotate with the body so that the moment of inertia is constant in time. In this body-xed coordinate system,
the conservation of angular momentum is
= d ([I]{}) = AppliedM oments
H
dt
2

(7)

Since we have chosen to work in a rotating coordinate system so that

d
dt I

= 0, we must pay the price, applying

Coriolis theorem to obtain the time derivative of the angular velocity vector in the rotating coordinate system
= d H + H,
H
dt

(8)

resulting in the Euler equations expressed in the x, y, z coordinate system moving with the body. In
general, we must rotate with the total angular velocity of the body, so that the governing equation for the
conservation of angular momentum become, with = .
Mx

= H x Hy z + Hz y

(9)

My

= H y Hz x + Hx z

(10)

Mz

= H z Hx y + Hy x

(11)

where is the rotational angular velocity of our axis system. In this case because of the symmetry of the
body, we are free to choose to allow the spin velocity to rotate relative to our body-xed axes system and
fasten our axis system to and . That is, we follow the motion of the top in and but allow it to spin
with angular velocity relative to our coordinate system rotating with angular velocity . So that

x =

(12)

y = sin

(13)

z = cos.

(14)

For this choice of coordinate system, we have essentially performed a rotation in and only, leading to
the geometry shown.

In this coordinate system, since the rotation did not occur, the angular velocity of the body is
x =

(15)

y =

sin

(16)

+ .
z = cos

(17)

For this choice of coordinate system, we have


Hx =

I0

(18)

Hy =

I0 sin

(19)

Hz = I( cos + )

(20)

resulting in
H x =

I0

(21)

H y =

cos
I0 (sin
+

(22)

H z = I( cos sin + )

(23)

and H in components is
(I I0 ) cos sin 2 + I sin

(24)

(I0 I) cos I

(25)

0.

(26)

Note that the z component is zero. Since the axes chosen are principal axes, the nal form of Eulers
equations becomes,

Mx

= I0 ( 2 sin cos ) + I sin ( cos + )

(27)

My

= I0 ( sin + 2 cos ) I( cos + )

(28)

Mz

= I( + cos sin )

(29)

For a top, Mx = M gzG sin, My = 0 and Mz = 0. These equations are unsteady and non-linear. We can
gain insight by examining the character of some special solutions and constants of the motion.

Steady Precession: Gyroscopic Motion


We now consider the steady precession of a top about the Z axis. In terms of the variables we have dened,
the top rotates with spin velocity about its principal axis, and precesses with angular velocity while

maintaining a constant angle with the vertical axis Z. We then have


= constant = 0

= 0

(30)

= constant = 0

= = 0

(31)

= 0

(32)

= 0

For a steady precession of a top, Eulers equations reduce to

sin(I(cos
+ ) I0 cos)
= Mx = M gsinzG

(33)

Note that sin cancels, resulting in


I (I0 I) 2 cos = M gzG

(34)

In the usual case for tops or gyroscopes, we have >> so that 2 may be ignored. Therefore, for steady
precession, the relationship between the precession angular velocity and the spin angular velocity is
= M gzG /(I )

(35)

where I = Izz , the moment of inertia of the gyroscope about its spin axis. This result is also true if = 0. For
this motion, the angular momentum vector is not aligned with the Z axis as for free-body motion, but is in
the plane of z, Z, and rotates around the Z axis according to the applied external moment which is constant
and in the x direction. In the limit as >> the angular momentum vector is essentially along the axis of
rotation, z, with a slight component due to which can be calculated after has been determined. Some

attempt to sketch this is shown in the gure, but even this small correction is an approximation for >> .
For >> , this same result can be obtained by less complex approach, but it is important to realize that
this is an approximation. For the spin velocity much greater than the precessional velocity, we may take
5

the angular momentum vector as directed along the spin or z axis, H = Ik. Then the applied moment
will precess the vector with rotation rate H. Since = K = K, (where K is a unit vector in the Z
direction), this give a result for in agreement with the previous result
= M gzG /(I)

(36)

Since dropped out of the previous equation, for = 0, we again have = M gzG /(I).

The limit >> is the gyroscopic limit where the device behaves as a gyroscope rather than as the more

general case of a top. The dierence is that, for a gyroscope, is larger than any other rotation rate in the
system, such as the angular velocity of an aircraft or spacecraft. This makes the gyroscope a useful basis for
many instruments. We shall return to this issue.

Unsteady Precession of a Top: Integrals of Motion


For the general case of a top in a gravitational eld, we have Eulers Equations. In the most general case,
we will have spin , precession and nutation , all varying with time.
Mx

= I0 ( 2 sin cos ) + I sin (cos


+ ) = M gzG sin

(37)

My

= I0 ( sin + 2 cos ) I( cos + ) = 0

(38)

Mz

= I( + cos sin ) = 0

(39)

The consequences of My = 0 and Mz = 0 provide two quantities that must be constant/conserved during
the motion. (We will later show that these results can also be obtained by applying Lagranges equation to
this system.)

d/dt(I0 sin2 + I cos ( + cos))


=

dp
dt

=0

(40)

d/dt(I( + cos )) =

dp
dt

=0

(41)

It can easily be seen by inspection that dierentiating p and setting it equal to 0 yields the equation Mz = 0.
The equation from

dp
dt

is more complex; to obtain My requires combinations of


6

dp
dt

and

dp
dt .

But the conclusion is that p and p are constants of the motion; we will show later that these can be
considered generalized momenta, derived by application of Lagranges equation.
Therefore, the spin angular velocity and the precession angular velocity are instantaneously related to the
nutation angle through
p p cos
=
I0 sin2
p
p p cos

=
Isin2
I0

(42)
(43)

We may therefore write the remaining governing equation as


+ (

p
I0

p
p
I cos)( I
sin3

p cos
)
I0

gM zG
=0
I0

(44)

This reduces the problem to the motion of a single variable ; once (t) has been determined, the constancy
of p /I0 and p /I during the motion, give the relations between the nutation angle (t), the spin angle (t),
and precession angle (t). For this more complex motion, all three angles change with time, and the tip of
the top traces out a motion, inscribed on the surface of a sphere for visualization, as shown in the gure.
These various motions are referred to as unidirectional precession, looping precession and cuspidal motion.

For more discussion of these solutions see J.B Marion and S.T. Thornton, Classical Dynamics, Chapter 11.

Spinning Top by Lagranges Equation


The constancy of two momenta obtained by application of Eulers equation can be found perhaps more
directly by application of Lagranges equation. We write the kinetic energy of a spinning top as
T = 1/2I0 (x2 + y2 ) + 1/2Iz2

(45)

We use the full form of the angular velocities in the Euler system moving with the spinning top and obtain
x2 = ( sin sin + cos )2

(46)

y2 = (sin
cos sin )2

(47)

2
z2 = (cos
+ )

(48)

resulting in
x2 + y2 = 2 sin2 + 2

(49)

z2 = ( cos + )2 .

(50)

T = 1/2I0 ( 2 sin2 + 2 ) + 1/2I( cos + )2

(51)

Therefore the kinetic energy is

For a top, the potential energy is


V = M gzG cos

(52)

L = 1/2I0 ( 2 sin2 + 2 ) + 1/2I( cos + )2 M gzG cos .

(53)

so that the Lagrangian is

We consider , and to be our generalized coordinates, qi . And we move forward applying Lagranges
equation, d/dt(L/qi ) V /qi = 0. However, since V /qi = 0 for both qi = and qi = , we conclude
that
L
= (I0 sin2 + I cos2 ) + I cos = constant = p

L
= I( + cos ) = constant = p

(54)
(55)

We dene these groupings which must be constant in time as generalized momenta, consistent with our
earlier exposition of Lagranges equations. These are of course the same constants identied using Eulers
equations in equations (38-39). Lagranges equations win! The nal governing equation is obtained from

d L
L

=0
(56)
dt

resulting in
+ (

p
I0

p
p
I cos)( I
sin3

p cos
)
I0

gM zG
= 0,
I0

(57)

in agreement with equation (44).


We also observe that total energy is a constant of the motion.
E = 1/2 I0 ( 2 sin2 + theta2 ) + 1/2I0 cos + )2 + M HzG cos

(58)

For more discussion of these solutions see J.B Marion and S.T. Thornton, Classical Dynamics, Chapter 11.

Extreme Tops: Gyroscopes


Considerable simplication and practical application results if the angular velocity is an order of magnitude
larger than angular velocity associated with the motions of a system of interest, such as an aircraft or a
8

spacecraft. In this case, the gyroscope can be used as an instrument to measure some quantity of interest in

a vehicle.

In order to use a gyroscope as an instrument, we rst have to consider how it is mounted and axed to the

vehicle. The classic mount is called a gimbal. It permits free rotation of the gyroscope about its own axis,

and depending on how complex the gimbal system is, free rotation about other axis as well.

Consider the gyroscope shown in the gure. It is free to spin about its axis with angular velocity . It is
attached to a gimbal which is free to rotate about the support as sketched. The entire apparatus is mounted
on a turntable rotating with angular velocity , where >> . If we can measure the orientation of the
axis of rotation of the disk relative to the turntable mount, we could possibly use this as an instrument.
What does it measure?
Consider that the initial angular momentum vector of the disk is in the horizontal plane. The axis of the
disk is free to rotate about the horizontal axis through an angle . As the turntable rotates with angular
velocity , the axis of the disk will rotate in its only free direction to keep its rate of change of angular
momentum zero, since it can experience no moments/torques. There are two source of changes in angular
momentum, the unsteady component if changes with time; and the component of the angular momentum
of the disk in the horizontal plane that is precessed by the turntable rotation .

where I0 is the moment of inertia of the disk about the axis perpendicular to
The unsteady term is I0 ,
its spin axis; the vector is in the horizontal direction. The horizontal component of the angular momentum
of the disk is Isin, where I is the moment of inertia of the disk about its axis of rotation. The rate of

change is also in the horizonal direction, perpendicular to and of magnitude Isin. The sum of these
two terms, which must be zero, is
I0 + Isin = 0

(59)

This is the familiar equation for the oscillation of a pendulum. If there is a small amount of friction in the
support, the system will settle at = 0 and will indicate the direction of rotation of the turntable axis.
This is perhaps not too interesting, since we already know the direction of rotation of the turntable. But if
we have this device on a rotating body, such as the earth, it will again indicate the direction of the axis of
rotation of the body, in this case the earth. We call this direction North. Thus, this application of the
gyroscope could be used as a compass, often called a gyrocompass.
ADDITIONAL READING
J.L. Meriam and L.G. Kraige, Engineering Mechanics, DYNAMICS, 5th Edition
7/9
J.B Marion and S.T. Thornton, Classical Dynamics, Chapter 11

W.T. Thompson, Introduction to Space Dynamics, Chapter 5

J. H. Ginsberg, Advanced Engineering Dynamics, Second Edition, Chapter 8

D. Kleppner, R.J. Kolenkow, An introduction to Mechanics,Chapter 7

10

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

Inertial Instruments and Inertial Navigation


Gimbals
Gimbals are essentially hinges that allow freedom of rotation about one axis.
Gimbals often have superb bearings and motors to help achieve virtually
frictionless behavior. Sensors in the bearings provide measurements of
gimbal angles. Three gimbals allow freedom of rotation of a vehicle about
three axes while a central platform remains stationary with respect to inertial
space.
Gvros
A gyro is a spinning mass with relatively large angular momentum. We
know that the rate of change of angular momentum is equal to the applied
moment.

-A -H = M-

6Itr

If no torque is applied then the angular momentum vector remains stationary


with respect to inertial space. Gimbals allow a vehicle to rotate ii-eely about
a gyro so the gyro spin axis can provide a single axis direction that is
stationary with respect to inertial space.
Restraining a gyro about an axis perpendicular to the angular momentum
vector provides a means for measuring angular velocity with respect to
inertial space. This device is called a rate gyro and is a common sensor for
aiding in rate stabilization of vehicles (e.g., the D in a PD controller).
Inertial Platforms
A gyro mounted on a platform can be used as a sensor in a feedback loop to
stabilize the platform with respect to inertial space. This is called an
inertially stabilized platform. A S we will see, the inertially stabilized
platform is an essential element of inertial navigation.

Applying torque to the gyro causes its spin vector (i.e., angular momentum
vector) to move with respect to inertial space. Thus the inertially stabilized
platform can be reoriented with respect to inertial space.
Accelerometers
A second important inertial sensor is the accelerometer. A simplified
diagram of an accelerometer is as follows&--.
... .

~1 d e h i c \ e

C \e
~ VOwr4~r

cc.se.

Where
m=test mass
d=displacement of the vehicle from an inertially fixed point
x=displacement of the test mass from its rest point
x+d=displacement of the test mass from the inertially fixed point

x,=transducer output signal


thus

So the system differential equation is

which is a second order LTI system. Vehicle acceleration as the input and
the output is the negative of indicated test mass displacement times klm.

Note in particular that if the vehicle acceleration is constant then the steady
state output is constant, thus producing an indication of that acceleration.
The undamped natural frequency and damping ratio of the accelerometer
are
bJwT

jg'

s=-

where the parameters c and k are controlled by the manufacturer. Typical


.
values are

The following figure illustrates the response of such a system to a very short
one "g" pulse of vehicle acceleration that is 20 milliseconds in duration.
-

--

Acceleration Pulse Input and Accelerometer Response

-0.2

time (sac)
-~

-.

~-

----

~p

~~

Commonly vehicle velocity is desired so the accelerometer output is


integrated over time

The following figure illustrates the vehicle velocity produced by the


acceleration pulse shown above, compared with the time integral of the
accelerometer output.
- -

- --

- --

Vehicle Velocity and the Integral of Accelerometer Output

0
0

0.005

0.01

0.015

~~
~
~~
~p~

-----

0.02
time (sec)
~

0.025

0.03
-

~p~

0.035
~

0.04
-

p~

Note that, except for a small delay, the integral of the accelerometer output
is a very good representation of vehicle velocity.
Spacecraft System Applications of Inertial Systems
Space systems utilize inertially stabilized platforms in a number of ways
-provide a reference for stabilizing and controlling vehicle attitude
-stabilize sensors and point them in desired directions
-provide a stable reference for estimating changes in vehicle velocity

Consider the boost of a spacecraft ftom low earth orbit (LEO) onto a
trajectory to the moon. Rocket motors must increase the vehicle velocity by
the order of thousands of meters per second, but with a level of precision
that is on the order of only meters per second. In other words the velocity
change must be accomplished so that the error is only about 0.1% of the
required boost in velocity.

The boost velocity is achieved using an inertial platform as the primary


sensor of the vehicle acceleration. The inertial platform looks like this2 d~te\erovht+<*

A*

The three gyros stabilize the platform orientation so that it is stationary with
respect to inertial space. Each gyro is responsible for stabilization about one
axis. The stable platform holds the accelerometers fixed in space so that
they can sense acceleration. The three components of acceleration are then
integrated to estimate the change in vehicle velocity vector.
The following diagram illustrates the essential elements of the feedback
system used to accomplish this function.

Note the inner loop that controls vehicle attitude and the outer loop that
controls the velocity change imparted by the rocket motors. The inner loop
utilizes both attitude and angular velocity measurements, from the inertial
platform gyros, to stabilize and guide the vehicle. The outer loop utilizes
integrated accelerometer outputs to achieve the desired velocity change.

Aircraft System A ~ ~ l i c a t i o n
ofs Inertial Svstems
Modem aircraft systems use inertial platforms to implement combined
inertial and GPS navigation systems. Typically the inertial system
accomplishes the navigation function itself, especially over short time
intervals. However, gyro and accelerometer drift and bias errors tend to
degrade performance over time, so a GPS receiver serves to periodically
correct these errors. In effect the inertial system serves as the short term
(high bandwidth) sensor and the GPS serves as the long term (low
bandwidth) sensor.
Often an inertial platform is used to track the local north, east and vertical
(down) directions at the vehicle location. The following diagram illustrates
this configuration

In effect the inertial platform provides a replica of the local horizon within
the aircraft and serves as the reference for the horizontal situation indicator
on the aircraft instrument panel. The inertial platform gimbals allow the
inertial platform to remain stable, with respect to the local vertical, even as
the vehicle may perform violent angular maneuvers about the stable inertial
platform.
In order to maintain the local vertical the platform must rotate at an angular
velocity that matches the sum of earth rate and the rate at which the vehicle
moves over the Earth. For example, in order to match Earth rate plus the
rate at which the vehicle moves in the east or west directions, the system
must rotate about a vector pointing in the direction of the Earth's axis of
rotation. Thus, if the vehicle is at some latitude h then the north, and down
gyros must be torqued so their axes rotate at a rate equal to the sum of Earth
rate plus the rate of change of vehicle longitude, as shown in the following
diagram.

Similarly, as the vehicle moves in the north or south directions the east gyro
must be torqued so that the platform rotates at the rate of change of latitude.
Earth rate is well known to very high accuracy. However the inertial system
must keep track of both position and velocity so that both the local vertical
can be maintained and that latitude and longitude are available as system
outputs to the vehicle crew. Accelerometers mounted on the inertial
platform provide measurements of north, east and down accelerations.
These are integrated to maintain north, east and down velocities, which in
turn are integrated to maintain latitude, longitude and altitude of the aircraft.
For example, north acceleration is integrated to maintain north velocity,
which is divided by the local Earth radius to obtain latitude rate, which is
again integrated to maintain latitude, as shown in the following diagram.

In addition, the attitude rate indication is used to torque the north and down
gyros as indicated above. A similar function is performed on the east
accelerometer outputs so as to maintain longitude.

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

SOME DYNAMICS AND CONTROL CHALLENGES


THAT OCCURED DURING THE APOLLO PROJECT
William S. Widnall, ScD
Formerly Director, Control and Flight Dynamics,
M.I.T. Instrumentation Laboratory
INTRODUCTION
President Kennedy set the goal to put a man on the moon
and return him safely before the end of the 60's decade.
The M.I.T. Instrumentation Lab (now called the Draper Lab,
in honor of its founder) was awarded by NASA the first
prime contract -- a contract to develop the navigation,
guidance, and control system for both the Command
Module and the Lunar Module of the Apollo Spacecraft.

Pitch axis y

Roll axis x

Yaw
axis z

Apollo Spacecraft in Docked Configuration

Lunar Landing Mission Phases

Command Module Navigation Guidance & Control System Components

Lunar Module Navigation Guidance & Control System Components

Apollo Development Flights

A DYNAMICS CHALLENGE -- APOLLO BARBEQUE MODE


Requirement: For long periods of coasting flight, provide passive thermal control by implementing a
"barbeque mode" -- a very slow rotation (one revolution per ten minutes or slower) about the
spacecraft roll axis and with the desired roll axis direction fixed in space.
Design provided: An attitude control mode that used the reaction control jets to establish the initial
desired angular velocity, and then as needed to maintain the spacecraft attitude close to the rotating
desired attitude associated with the desired constant angular velocity vector along the roll axis with
a fixed direction in space.
In-Flight Performance: There was a slight dynamic imbalance because the moment of inertia
principal axes were not precisely aligned with the spacecraft axes. Active torquing was required to
maintain the spacecraft attitude close to the desired rotating attitude. While the fuel consumption
was thought to be acceptable, nevertheless the astronauts complained that the banging of a jet
every minute or so made it impossible for them to sleep. NASA asked us at M.I.T. what to do?
First suggestion: Our first suggestion was to turn off the active automatic attitude control (no jet
firings) after the initial desired angular velocity was established. It was hoped that the subsequent
free-body motion would be reasonably close to the desired spinning motion. The astronauts gave
this a try, but the subsequent wandering of the roll axis away from its desired azimuth and elevation
were deemed too large. What else to try?
Second suggestion: Bill Widnall suggested that what was needed was to get the angular momentum
vector much closer to the principal axis nearest to the roll axis and that this could be accomplished
by not using the active attitude control mode to establish a precise roll angular velocity but rather by
using the rotational hand controller to command a pure torque about the roll axis. The induced
angular momentum vector would be aligned with that torque impulse. This worked very well. In the
subsequent free-body motion the roll axis deviation away from its desired orientation was
acceptably small and the astronauts got their sleep.

Active torquing required to have spacecraft roll axis remain aligned


with desired angular velocity vector, which is fixed in space

H=M

wx

Pitch axis y

Roll axis
x

Yaw
axis z

Differential equations to simulate spacecraft free-body motion


Case 1 - Initially only roll angular velocity is non-zero (1 rev / 10 min)
Case 2 - Initial conditions are established by a roll axis torque impulse

degrees

30
20

azimuth

10

50

100

150

200

Time-minutes

-10
-20
-30

elevation

degrees

elevation 5
-5

5
-5

degrees

-10

Free-body motion of
spacecraft roll axis after
establishing angular velocity
solely along the roll axis

-15
-20

10

15

azimuth

20

initial roll axis


principal axis

fixed in space

11.4 deg
1.38 deg

body cone

space cone

Free-body motion with initial w parallel to roll axis is very


close to that predicted when the two largest principal
moments of inertia are equal

Free-body motion due to torque impulse applied about roll axis


initial roll axis and initial torque impulse

principal axis

1.38 deg

fixed in space

1.23 deg

.15 deg
path of roll
axis
body cone

space cone

degrees

elevation

elevation
50

100

150

200

2
-2

degrees

-4

Time-minutes

azimuth
-2

azimuth

-1

-1
2

-2
50

100

150

200

-2
-4

Time-minutes

Free-body motion of spacecraft roll axis


after applying torque impulse about the roll axis

elevation
5

degrees

azimuth

2
1

-2

-5

-1

1
-1
-2

10

15

20

degrees

-5
-10
-15
-20
Comparison of free-body motions of spacecraft roll axis for the two cases:
angular velocity initial condition versus torque impulse initial condition

A "BACKUP" CONTROL MODE THAT HELPED SAVE LIVES -USING THE LUNAR MODULE TO PUSH THE COMMAND MODULE
Requirement: Provide a capability for the Lunar Module to push the Command and Service Module, in case the
CSM were to become disabled.
Challenges: The LM was not specifically designed to accommodate this requirement. When in the docked
configuration, the LM reaction control jets had exhaust impingment problems: To use the jets that exhausted
upward against the attached command module was not acceptable. The downward exhausting jets were known
to impinge against the LM descent stage and this would produce significant adverse torque because of the more
distant center of mass location.
Design approach: During powered flight use the thrusting LM descent-stage engine, rather than the reaction
control jets, to control the spacecraft pitch and yaw. There was an engine gimballing capability that could
change the engine thrusting angles in the pitch and yaw planes at a very slow rate of 0.2 deg / sec. The initial
intended use of this capability was to null out any pitch and yaw torques so that reaction jets would not have to
be used continually to balance the bias torques. Bill Widnall proposed that it might be possible to do spacecraft
pitch and yaw attitude control using this slow gimballing capability of the descent engine. Because the gimbal
rates were so slow, Widnall sought and successfully derived the minimum time optimal control law for the thirdorder dynamic systems in the two (pitch and yaw) planes. Simulation results indicated that this LM nonlinear
minimum-time thrust-vector control law would be able to control the docked configuration in pitch and yaw
without assistance from the reaction control jets.
Apollo 13: On the way to the moon during the Apollo 13 mission, an explosion in the Service Module disabled
the Service Module including its main engine. The lunar landing goal was aborted and the challenge became,
could we get the astronauts home? Many things had to work, including the using of the undamaged LM to push
the spacecraft when trajectory corrections were needed to maintain the free-return-to-earth trajectory.
Fortunately the M.I.T. team had provided the backup control capability that now was needed. The astronauts
were returned safely to earth

Descent Configuration of the Lunar Module: Reaction Control


jet exhaust impingement weakens the applied torque

With the longer moment arm in the docked configuration


the impingement force exerts a significant adverse torque

q
G
F
d
+0.2 deg/sec
0
Control variable d =
-0.2 deg/sec
Thrust vector control of the docked configuration
using the LM descent engine

The dynamics in the pitch or yaw plane is that of a tripleintegral plant.

Example of minimum time attitude control:


Response to an initial thrust misalignment at ignition

The minimum-time thrust-vector control law in the LM


digital autopilot

Information flow in the LM digital autopilot


during descent engine powered flight

MIT OpenCourseWare
http://ocw.mit.edu

16.07 Dynamics
Fall 2009

For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

You might also like