You are on page 1of 118

University of Glasgow

Lecture Notes for the Courses

Spaceflight Dynamics 5 (M.En.)


Spaceflight Dynamics 2 (M.Sc.)
Academic Year 2009 - 10

Giulio Avanzini
Dipartimento di Ingegneria Aeronautica e Spaziale
Politecnico di Torino
e-mail: giulio.avanzini@polito.it

Contents
I

Attitude Dynamics and Control

1 Rigid Body dynamics


1.1 Frames of reference and transformation matrices . . . . . . . . . . . . . .
1.2 Eulers angles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2.1 Building the coordinate transformation matrix from elementary rotations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2.2 Angular velocity and the evolution of Eulers angles . . . . . . . .
1.2.3 The quaternions . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2.4 Evolution of the quaternions . . . . . . . . . . . . . . . . . . . . .
1.2.5 Quaternions vs Euler angles . . . . . . . . . . . . . . . . . . . . .
1.2.6 Other attitude representations . . . . . . . . . . . . . . . . . . . .
1.3 Time derivative of vector quantities . . . . . . . . . . . . . . . . . . . . .
1.4 Eulers equations of motion of a rigid body . . . . . . . . . . . . . . . . .
1.4.1 The inertia tensor . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.4.2 Rotational kinetic energy . . . . . . . . . . . . . . . . . . . . . . .
1.4.3 Eulers equation of motion . . . . . . . . . . . . . . . . . . . . . .
1.4.4 Conservation of angular momentum . . . . . . . . . . . . . . . . .
1.4.5 Conservation of kinetic energy . . . . . . . . . . . . . . . . . . . .
1.5 Generalised Euler equation . . . . . . . . . . . . . . . . . . . . . . . . . .
1.5.1 Derivation of the generalised form of Euler equation . . . . . . . .
1.5.2 Use of the generalised form of Euler equation . . . . . . . . . . .

.
.

3
3
5

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

7
11
12
16
16
16
17
18
18
21
22
22
23
23
23
25

2 Dynamics of Spinning Spacecraft


2.1 Torquefree motion of axisymmetric satellites . . . . . . . . .
2.2 Torquefree motion of triinertial satellites . . . . . . . . . . .
2.2.1 Drawing the polhode curves . . . . . . . . . . . . . . .
2.2.2 Stability of torquefree motion about principal axes . .
2.3 Nutation damping . . . . . . . . . . . . . . . . . . . . . . . .
2.3.1 Effects of energy dissipation on axisymmetric satellites
2.3.2 More accurate models . . . . . . . . . . . . . . . . . .
2.4 Attitude Maneuvers of a spinning satellite . . . . . . . . . . .

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

27
27
33
34
36
37
37
39
40

3 Dualspin satellites
3.1 Mathematical model of a gyrostat . . .
3.1.1 Angular momentum formulation
3.1.2 Angular velocity formulation . .
3.2 Simplified models . . . . . . . . . . . .
3.2.1 The Kelvin gyrostat . . . . . .
3.2.2 The apparent gyrostat . . . . .

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

45
47
47
48
49
49
50

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

ii

CONTENTS
3.3

Stability of axial gyrostat

. . . . . . . . . . . . . . . . . . . . . . . . . . . 50

4 Active Stabilisation and Control of Spacecraft


4.1 Actuators . . . . . . . . . . . . . . . . . . . . . . .
4.2 Linear model of rigid satellite attitude motion . . .
4.3 Linear model of gyrostat attitude motion . . . . . .
4.4 Use of thrusters for attitude control . . . . . . . . .
4.4.1 Single axis slews (open loop) . . . . . . . . .
4.4.2 Closedloop control . . . . . . . . . . . . . .
4.4.3 Fine pointing control . . . . . . . . . . . . .
4.5 Momentum exchange devices for attitude control . .
4.5.1 Openloop control with RWs . . . . . . . .
4.5.2 Sizing a reaction wheel for single axis slews .
4.5.3 Closedloop control with RWs for single axis
4.5.4 Bias torque and reaction wheel saturation .

. . . .
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .
slews
. . . .

5 Environmental torques and other disturbances


5.1 Environmental torques . . . . . . . . . . . . . . . . . .
5.2 Internal disturbances . . . . . . . . . . . . . . . . . . .
5.3 Gravitygradient stabilization . . . . . . . . . . . . . .
5.3.1 Origin of the gravitygradient torque . . . . . .
5.3.2 Pitch component of the gravitygradient torque
5.3.3 Attitude motion . . . . . . . . . . . . . . . . . .

II

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

53
53
54
55
56
56
59
65
69
70
72
73
73

.
.
.
.
.
.

75
75
76
77
77
78
80

Advanced Orbital Dynamics: Orbit Control

6 Basic concepts
6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . .
6.2 Keplerian motion: a review . . . . . . . . . . . . . .
6.2.1 Birth of Astrodynamics: Keplers Laws . . . .
6.2.2 Newtons Laws of Motion . . . . . . . . . . .
6.2.3 Newtons Law of Universal Gravitation . . . .
6.2.4 Equation of Motion in the Two-Body Problem
6.2.5 Potential Energy . . . . . . . . . . . . . . . .
6.2.6 Constants of the Motion . . . . . . . . . . . .
6.2.7 Trajectory Equation . . . . . . . . . . . . . .
6.2.8 Relating Energy and Semi-major Axis . . . .
6.2.9 Elliptical Orbits . . . . . . . . . . . . . . . . .
6.3 Three-Dimensional Analysis of Motion . . . . . . . .
6.3.1 Nonrotating Frames . . . . . . . . . . . . . .
6.3.2 Classical Orbital Elements . . . . . . . . . . .
6.3.3 Modified Equinoctial Orbital Elements . . . .
6.3.4 Determining the orbital elements . . . . . . .
6.3.5 Orbit propagation: the ideal case . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

83
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

85
85
87
87
88
89
89
90
90
91
92
93
95
95
96
97
97
98

CONTENTS
7 Orbit Trim Manoeuvres
7.1 Variational Equations . .
7.2 Perigee Raise Manoeuvre
7.3 Plane Changes . . . . .
7.4 Finite Burn Manoeuvres

iii

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

101
. 101
. 103
. 105
. 105

8 Orbit Control
107
8.1 Air Drag Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
8.2 Geostationary Orbit Control . . . . . . . . . . . . . . . . . . . . . . . . . . 110

Part I
Attitude Dynamics and Control

Chapter 1
Rigid Body dynamics
In order to describe the attitude of a rigid body and to determine its evolution as a
function of its initial angular velocity and applied torques, Eulers angles and Eulers
equations of motion need to be introduced. The transformation matrix between different
reference frames will be recalled and the concept of inertia tensor will also be briefly
discussed.

1.1

Frames of reference and transformation matrices

Assuming that a satellite is a rigid body is a reasonable initial model for attitude dynamics and control. However, in practice, this assumption can only be used as a first
approximation. For satellites with large deployable solar arrays the structure can be
quite flexible. The elastic modes in the structure can be excited through attitude control thrusters firings. This leads to vibrations which reduce the pointing accuracy of the
payload. In addition, fuel consumption and fuel slosh in propellant tanks can cause the
inertia properties of the satellite to be time varying, leading to a more complex control
problem. But if we assume that our spacecraft is a rigid body, we can attach to it a body
2, e
3 ). The position of FB with respect
frame, FB , described by a set of unit vectors (
e1 , e
1, E
2, E
3 ), completely
to an inertial reference frame FI , identified by the unit vectors (E
describes the attitude of our spacecraft.
Assuming that ~v is a vector quantity, it is possible to write it as

or, equivalently,

~v = x
e1 + y
e2 + z
e3

3
~v = X E 1 + Y E 2 + Z E

The column vectors v B = (x, y, z)T and v I = (X, Y, Z)T provide the component representations of the same vector quantity ~v in the reference frames FB and FI , respectively.
iI = (e1,i , e2,i , e3,i )T of the ith unit vector e
i in
If we now consider the components e
FI , that is
1 + e2,i E
2 + e3,i E
3
i = e1,i E
e
we can write
~v = x
e1 + y
e2 + z
e3

2 + e3,1 E
3) +
= x(e1,1 E 1 + e2,1 E
1 + e2,2 E
2 + e3,2 E
3) +
+ y(e1,2 E
1 + e2,3 E
2 + e3,3 E
3)
+ z(e1,3 E
3

G. Avanzini, Spaceflight Dynamics 1. Rigid Body dynamics

1 +
~v = (e1,1 x + e1,2 y + e1,3 z)E
2 +
+ (e2,1 x + e2,2 y + e2,3 z)E
3
+ (e3,1 x + e3,2 y + e3,3 z)E
This means that the components of ~v in FI can be expressed as a function of those in
FB as follows:
X=e1,1 x + e1,2 y + e1,3 z
Y =e2,1 x + e2,2 y + e2,3 z
Z=e3,1 x + e3,2 y + e3,3 z
or, in compact matrix form,
v I = LIB v B
where the transformation matrix LIB is given by



e1,1 e1,2 e1,3
.. ..

1I .
LIB = e2,1 e2,2 e2,3 = e
e2I .
e3I
e3,1 e3,2 e3,3

i as expressed in FI . Any matrix


LIB is made up by the components of the unit vectors e
made up by mutually orthogonal row or column unit vectors is an orthogonal matrix and
is characterized by several properties, among which we only recall that:
the inverse of an orthogonal matrix L is given by its transpose: L1 = LT ;
the determinant of an orthogonal matrix is det(L) = 1 (and that of a rotation
matrix is 1);
if L1 and L2 are orthogonal matrices, their product L1 L2 is an orthogonal matrix.
Thanks to the first property, it is possible to write the inverse coordinate transformation as
v B = LBI v I = (LIB )1 v I = (LIB )T v I
which means that it is also

LBI

T1I
e
...
T2I
e
...
T3I
e

As an exercise, demonstrate that the dual relations

LBI

also hold.




.
.

1 ..E
2 ..E
3
= E
;
L
=

IB
B
B
B

T
E
1B
...
T
E
2B
...
T
E
3B

G. Avanzini, Spaceflight Dynamics 1. Rigid Body dynamics

1.2

Eulers angles

It is possible to use the coordinate transformation matrix LBI to describe the attitude
i of the body frame attached to it, coming
of the spacecraft through the unit vectors e
out with a total of 9 parameters. As a matter of fact, these 9 parameters are not free to
vary at will, inasmuch as they must satisfy 6 constraints, expressed by the orthonormality
condition, that is

0 if i 6= j
i e
j = i,j =
e
1 if i = j
Roughly speaking, only 9 6 = 3 parameters should be sufficient to describe the attitude
of FB w.r.t. FI .
One of the set of three parameters most widely used to describe the attitude of a rigid
body (or equivalently the attitude of the body frame attached to it) w.r.t. a fixed frame
are the Eulers angles, a sequence of three rotations that take the fixed frame and make
it coincide with the body frame. The original sequence of rotations proposed by Euler to
superimpose FI onto FB is the sequences 3-1-3:
1. the first rotation is about the third axis of the initial frame, that is E3 , in our case,
1 to the direction e
1 perpendicular to the plane determined
and takes the first axis E
3 and e
2 is rotated onto e
3; E
2 ; the rotation angle is called
by the unit vectors E
precession angle ;
1 ,
2. the second rotation is about the first axis transformed after the first rotation, e

3 into the position of e


3; e
2 is moved onto e
2 ; the rotation angle
and takes the axis e
is called nutation angle ;
3 and brings e
1 = e
1 and e
2 to their final
3. the third and final rotation is about e
1 and e
2 , respectively; the rotation angle is called spin angle .
positions, e
The three angles, representing the amplitude of the three, successive rotations , , ,
respectively about the third, the first, and again the third axis, can be used to represent
the attitude of the frame FB : The nutation angle represents the inclination of the third
3 ; The precession angle represents the angle between
3 w.r.t. the local vertical E
body axis e

the first inertial axis E 1 and the line of the nodes , i.e. the intersection between the
3 ; The spin angle is the rotation about the third body
3 and E
planes perpendicular to e
axis.
The transformation matrix LBI can be expressed as a function of these three angles,
in terms of three elementary rotation matrices, as will be derived in the sequel.
Other sequences
It must be remembered that the sequence of rotations here described is not the only
possible choice for rotating FI onto FB . Many other sequences are available and equally
useful. In atmospheric flight mechanics the most widely used sequence of rotations is the
321, also known as the Bryants angles.
3 , and its amplitude is called
In this case the first rotation is about the third axis, E
2 is the pitch angle, , and
yaw angle . The second rotation about the second axis e
1 is the roll angle,
takes the first axis onto its final position. The third rotation about e
. This set of angles is used also in space flight dynamics, to describe the attitude of a

G. Avanzini, Spaceflight Dynamics 1. Rigid Body dynamics

3
E

3
3 E
e

2
e

1
E

2
E
3
E

1 e
e
1

2
E

2
e

1
E

2
e

3 e
3
e
2
e

1
E

2
E

1
e
3
E

2
e

3
e

1
E

2
e
1
e

1
e

2
E

Figure 1.1: Eulers angles


spacecraft with respect to the Local Horizontal Local Vertical (LHLV) reference frame.
In many textbooks also this latter set of rotations is often referred to as Eulers angles,
and this fact may lead to some confusion.
As a final observation, the order of the rotation sequence is important: Rotations do
not commute! This means that the rotation sequence 123, performed with the same
angles about the same axis, will take the initial frame to another one. The rotation
sequence 123 is known as Cardan angles.
Singularity
There is another problem with the representation of rotations in a three dimensional space,
that is the singularity of all the descriptions in terms of three parameters. This means
that there will always be positions of the two frames that can be described in different
ways, once a particular sequence of rotations is chosen. As an example, if the original
Eulers angle sequence is employed, the case in which = 0 is singular, inasmuch as the
3 e
3 . This means
precession and spin rotations will be about the very same axis, E
that all the triplets (, 0, ) for which + is constant represent the same change of
reference frame.
Similarly, when the Bryants angles are used, the case = /2 is singular, as in this
case all the triplets (, /2, ) for which is constant will provide the same final
attitude for FB .
The problem of coordinate transformation singularity has some unpleasant mathemat-

G. Avanzini, Spaceflight Dynamics 1. Rigid Body dynamics

y
y

y1

v
x

x1

Figure 1.2: Planar rotation of amplitude .


ical consequence that will be underlined in the following paragraphs.

1.2.1

Building the coordinate transformation matrix from elementary rotations

Planar rotations
Consider the sketch of Fig. 1.2, where two planar reference frames F1 and F2 , with the
same origin O are represented. The angle , assumed positive for counterclockwise
rotations, allows one to identify univocally the position of the axes X2 Y2 of F2 w.r.t. the
frame F1 defined by the axes X1 and Y1 .
Given the components x1 and y1 of a vector ~v expressed in F1 , the components x2 e
y2 can be expressed as a function of the angle . The following relations can be easily
inferred from Fig. 1.2:
x2 = x1 cos() + y1 sin()
y2 = x1 sin() + y1 cos()
or, in matrix form,


x2
y2

cos() sin()
sin() cos()



x1
y1

This relation expresses the coordinate transformation that takes the component of a
vector quantity expressed in F1 into those of a reference frame F2 rotated w.r.t. F1 of an
angle .
In compact notation we can write
v 2 = L21 v 1 = R()v 1

G. Avanzini, Spaceflight Dynamics 1. Rigid Body dynamics

where the subscript near the vector indicates the frame in which the components of the
vector quantity are considered, the matrix L21 is the coordinate transformation matrix
from F1 to F2 that in the two dimensional case coincides with the elementary rotation
matrix R().
The inverse transformation from F2 to F1 is given by
v 1 = L12 v 2 = (R())1 v 2
Recalling the properties of orthogonal matrices, it is
L12 = (R())1 = (R())T = R()
Elementary rotations for the sequence 313
Each one of the Eulers rotations can be considered an elementary rotation about a given
axis, that remains unchanged during the transformation. It is still possible to apply
the relations derived for the planar case, adding a further equation that states that the
coordinate relative to the rotation axis does not vary.
The coordinate transformation during the first rotation is given by
x = X cos() + Y sin()
y = X sin() + Y cos()
z = Z
that, in matrix form, can be written as:

cos() sin() 0 X
x
y
Y
= sin() cos() 0

z
0
0
1
Z

In an analogous way it is possible to demonstrate that, during the second rotation


1 , the coordinate transformation is given by
about e
x = x
y = y cos() + z sin()
z = y sin() + z cos()
that in matrix form becomes:


1
0
0
x
x

y
y
= 0 cos() sin()


z
0 sin() cos()
z

3 is represented by the transformation


Finally, the third rotation about e
x = x cos() + y sin()
y = x sin() + y cos()
z = z

or, in matrix form:

cos() sin() 0 x
x
y
= sin() cos() 0 y


z
z
0
0
1

G. Avanzini, Spaceflight Dynamics 1. Rigid Body dynamics

The three elementary rotation matrices of the Eulers


defined as

cos() sin() 0
1

R3 () = sin() cos() 0
; R1 () = 0
0
0
1
0

cos() sin() 0

R3 () = sin() cos() 0
0
0
1

sequence 313 can thus be

0
0
cos() sin() ;
sin() cos()

where the subscript near the rotation matrix symbol R indicates the axis around which
the rotation is performed, while the argument indicates the amplitude of the rotation.
Summing up...
When passing from the inertial frame FI to the body frame FB using Eulers sequence, the
coordinate transformation of vector quantities can be obtained combining in the correct
order the elementary rotation matrices, as follows:
v = R3 ()v I
v = R1 ()v
v B = R3 ()v
that is
v B = R3 ()R1 ()R3 ()v I
This means that
LBI = R3 ()R1 ()R3 ()
Performing the rowcolumn products, the following expression for LBI is obtained:

cos cos
sin cos sin

LBI =

sin cos cos


+ cos sin

sin sin

sin cos

cos

cos cos sin cos cos cos cos sin

sin cos
sin sin

sin sin

As the product of orthogonal matrices is an orthogonal matrix, the inverse of which is


equal to its transpose, the inverse coordinate transformation matrix LIB is simply given
by

cos cos
cos cos sin
sin cos sin
sin cos

LIB = LBI 1 = LBI T =


sin cos cos
+ cos sin

cos sin

cos cos cos


sin sin
cos sin

sin sin

sin cos

cos

10

G. Avanzini, Spaceflight Dynamics 1. Rigid Body dynamics

Elementary rotations for the sequence 321


It is left as an exercise to the reader the composition of elementary rotation matrices for
the sequence 321. Adopting the same notation used above, it is

cos() sin() 0
cos() 0 sin()
;
1
0
R3 () = sin() cos() 0 ; R2 () = 0
0
0
1
sin() 0 cos()

1
0
0
R1 () = 0 cos() sin()
0 sin() cos()
and the final result is given by

LBI

cos cos

cos sin

sin

sin sin cos sin sin sin sin cos

+ cos cos
=
cos sin

cos sin cos cos sin sin cos cos


+ sin sin

sin cos

A first consequence of the Eulers angle singularity


When = 0, the coordinate transformation matrix does not depend on and separately, but only on their sum. In such a case, it is

cos cos
sin cos 0

sin sin + cos sin


cos(
+
)
sin(
+
)
0


LBI =
cos sin cos cos 0 = sin( + ) cos( + ) 0

sin cos sin sin

0
0
1
0

As an exercise, demonstrate that LBI depends on only, when Bryants rotation


sequence is employed and = /2.
How to build elementary rotation matrices
There is a simple way to build mnemonically the elementary rotation matrices. The
matrices are 3 by 3. If a rotation of an angle about the i-th axis is being considered,
place 1 in position i, i, and fill the remaining elements of the ith row and ith column
with zeroes. All the other elements of the principal diagonal are cos and the last two
outside the diagonal are sin . The sin element above the row with the 1 must have a
minus sign. As an example, let us consider a rotation about the second axis (like in the
second rotation of the Bryants sequence). We start filling the matrix with 0s along the
second row and column, with a one in position 2,2:

0
0 1 0
0

G. Avanzini, Spaceflight Dynamics 1. Rigid Body dynamics

11

2
e

3
E

2
e

3
e

2
e

1
e
1
E

2
E
1
e

Figure 1.3: Angular velocity as a funciton of Eulers angle rates.


Then we fill the diagonal with cos :

cos 0

0
1
0

0 cos

and we put sin in the remaining places,

cos
0
sin

with a minus sign in the row above the 1:

0 sin

1
0
0 cos

This is R2 (). When a rotation about the first axis is considered, the 1 is on the first row
and apparently there is no row above it. But it is sufficient to cycle and start again from
the bottom: In this case the minus sign is on the sin in the third row.

1.2.2

Angular velocity and the evolution of Eulers angles

~ is given by
The angular velocity
~ = 1 e
1 + 2 e
2 + 3 e
3

but it is also (see Fig. 1.3)

3 +
E
e1 +
e3
~ =

3 in FB are given by the third column of the


The components of the unit vector E
1
matrix LBI , that is E 3B = (sin sin , cos sin , cos )T , while the components of e
are (cos , sin , 0)T . Thus
~ =

+
+
=
+
+

(sin
sin
e1 + cos sin
e2 + cos
e3 ) +

(cos
e1 sin
e2 ) +
e3

sin sin +
cos )
(
e1 +

( cos sin sin )


e2
cos + )
e3
(

12

G. Avanzini, Spaceflight Dynamics 1. Rigid Body dynamics

or, in matrix form



sin sin cos 0
1

2
= cos sin sin 0


cos
0
1
3

Inverting the 3 3 matrix, one obtains the law of evolution of Eulers angles as a
function of angular velocity components in body axis, that is


sin / sin
cos / sin 0 1

cos
sin
0 2
=

sin / tan cos / tan 1


3

or, in explicit form,

= (1 sin + 2 cos )/ sin

= 1 cos 2 sin

= (1 sin 2 cos )/ tan + 3


These equations can be integrated to obtain the evolution of the Euler angles, if the
angular velocity is known. But they also show an unpleasant feature of Eulers angle
singularity, that is the spin and precession rates go to infinity when approaches 0. This
fact has some serious consequences on the problem of attitude representation, inasmuch
as it is not possible to accept a set of attitude parameters the evolution of which cannot
always be described in an accurate way.
If the Bryants angles are used, the reader can demonstrate that
e +
e3
3 +
~ = E

2
so that



1
0
sin
1

2
= 0 cos cos sin


3
0 sin cos cos

Also demonstrate that the inverse relation is

= 1 + (2 sin + 3 cos ) tan


= 2 cos 3 sin
= (2 sin + 3 cos )/ cos
Again, in the neighborhood of the singular condition = /2 the rate of change of the
roll and yaw angles goes to infinity.

1.2.3

The quaternions

Eulers eigenaxis rotation theorem states that it is possible to rotate a fixed frame FI
that is fixed in both
onto any arbitrary frame FB with a simple rotation around an axis a
frames, called the Eulers rotation axis or eigenaxis, the direction cosines of which are
the same in the two considered frame.
A very simple algebraic demonstration of Eulers theorem can be obtained from the
following considerations:

G. Avanzini, Spaceflight Dynamics 1. Rigid Body dynamics

13

the eigenvalues of any (real) orthogonal matrix L have unit modulus; indicating
with H the Hermitian conjugate, which, for a real matrix is coincident with the
transpose, one has
H
H a (1 )a

La = a aH LT La = a
a=0

that for any nontrivial eigenvector a implies that


= 1 || = 1

(at least) one eigenvalue is = 1; any n n real matrix has at least one real
eigenvalue if n is an odd number, which means that a 3 3 orthogonal matrix must
have at least one eigenvalue which is 1 = 1. The other couple of eigenvalues will
be, in the most general case, complex conjugate numbers of unit modulus, which
can be cast in the form 2,3 = exp(i). The determinant is equal to the product
of the eigenvalues, which is one, for an orthogonal matrix, so that
1 2 3 = 1 1 = 1
The eigenvector relative to the first eigenvalue satisfies the relation
La = 1 a
This means that there is a direction a which is not changed under the action of transformation matrix L. If L represents a coordinate change, the vector a will be represented
by the same components in both the considered reference frames,
= a1 e
1 + a2 e
2 + a3 e
3
a
1 + a2 E
2 + a3 E
3
= a1 E
For this reason, the transformation that takes the initial frame onto the final one can be
.
considered as a single rotation about the Euler axis a
In order to express the coordinate transformation matrix LBI as a function of and
it is sufficient to consider the following sequence of rotations:1
a
1 onto a
, so that the new frame F is given by the unit vectors
1. take the unit vector E

, e
2, e
3 ; call this rotation R;
a
2. rotate both frames FI and F about the eigenaxis of the rotation angle ; because
of the definition of Euler rotation, FI goes onto FB , while F will rotate into a
, e
2 , e
3 ; this rotation is represented by the
new frame F given by the unit vectors a
elementary rotation matrix R1 ();
that takes F onto F has the
3. at this point it should be noted that the rotation R
B
=R
but it is performed in the opposite direction so that R
T.
same magnitude of R,
Summing up it is
1

T R1 ()R

LBI = R

This derivation is taken from B. Wie, Space Vehicle Dynamics and Control, AIAA Education Series,
Reston (VA), USA, 1998, Chap. 5, pp. 312315 and 318320.

14

G. Avanzini, Spaceflight Dynamics 1. Rigid Body dynamics

3
E
(a)

1
E

3
e

3
E
(b)

3
e

2
e

2
E

2
2 e
E

1
E

1
e
3
E

3
e
(c)

1
E

3
e

3
e

3
e
(d)

2
e
2
e

2
e
2
e

2
e

E2

1
e

where

3
e

1
e

a1 a2 a3
= R21 R22 R23
R
R31 R32 R33

Carrying out the calculations, one get, for the components Lij of the coordinate transformation matrix LBI the following expressions:
2
2
L11 = a21 + (R21
+ R31
) cos
L12 = a1 a2 + (R21 R22 + R31 R32 ) cos + (R21 R32 R31 R22 ) sin
L13 = a1 a3 + (R21 R23 + R31 R33 ) cos + (R21 R33 R31 R23 ) sin

L21 = a2 a1 + (R22 R21 + R32 R31 ) cos + (R22 R33 R32 R23 ) sin
2
2
L22 = a22 + (R22
+ R32
) cos
L23 = a2 a3 + (R22 R23 + R32 R33 ) cos + (R22 R33 R32 R23 ) sin
L31 = a3 a1 + (R23 R21 + R33 R31 ) cos + (R23 R31 R33 R21 ) sin

G. Avanzini, Spaceflight Dynamics 1. Rigid Body dynamics

15

L32 = a3 a2 + (R23 R22 + R33 R32 ) cos + (R23 R32 R33 R22 ) sin
2
2
L33 = a23 + (R23
+ R33
) cos
one gets
Taking into account the orthogonality conditions for R,
2
2
a21 + R21
+ R31
=1
2
2
2
a2 + R22 + R32 = 1
2
2
a23 + R23
+ R33
=1

2
2
R21
+ R31
= 1 a21
2
2
R22 + R32 = 1 a22
2
2
R23
+ R33
= 1 a23

a1 a2 + R21 R22 + R31 R32 = 0 R21 R22 + R31 R32 = a1 a2


a2 a3 + R22 R23 + R32 R33 = 0 R22 R23 + R32 R33 = a2 a3
a3 a1 + R21 R23 + R31 R33 = 0 R21 R23 + R31 R33 = a1 a2

while remembering that the first row of an orthogonal matrix is given by the cross product
of the second and the third ones, it is
a1 = R22 R33 R23 R32
a2 = R23 R31 R21 R33
a3 = R21 R32 R22 R31
Substituting these results into the expressions of the coefficients Lij the following
expression for LBI is obtained:

cos + a21 (1 cos )


a1 a2 (1 cos ) + a3 sin a1 a3 (1 cos ) a2 sin
cos + a22 (1 cos )
a2 a3 (1 cos ) + a1 sin
LBI = a2 a1 (1 cos ) a3 sin
a3 a1 (1 cos ) + a2 sin a3 a2 (1 cos ) a1 sin
cos + a23 (1 cos )
or, in compact matrix form

T sin A
LBI = cos 1 + (1 cos )
aa
where 1 is the 3 3 identity matrix. The tilde symbol, V , is the equivalent matrix form
of the cross product, that is, letting

0 v3 v2
0 v1
V = v3
v2 v1
0

it is v w = V w.

We now define the Euler parameters or quaternions as


q0
q1
q2
q3

=
=
=
=

cos(/2)
a1 sin(/2)
a2 sin(/2)
a3 sin(/2)

sin(/2) and rembering that cos = cos2 (/2)sin2 (/2) = q02 qq and
By letting q = a
sin = 2 cos(/2) sin(/2) = 2q0 sin(/2), it is easy to demonstrate that the coordinate
transformation matrix is given by

LBI = (q 2 q q)1 + 2qq T 2q0 Q


0

where the indicates again the cross product matrix equivalent

0 q3 q2
= q3
0 q1
Q
q2 q1
0

16

G. Avanzini, Spaceflight Dynamics 1. Rigid Body dynamics

1.2.4

Evolution of the quaternions

The evolution of the quaternions is described


represented in matrix form as2

0 1

q1

0
1
=

q2

2 2 3

q3
3 2
The equivalent matrix form is given by

by the set of linear differential equations,

2 3
q0

3 2
q1
0
1
q2

q3
1
0

1
q0 = q
2
1
q =
(q0 q)
2

1.2.5

Quaternions vs Euler angles

The quaternions allow for representing the attitude of a rigid body with several advantages
over Eulers angles, above all the absence of inherent geometric singularity. Moreover, the
linear equation to be integrated in time in order to determine their evolution as a function
of angular velocity components is less computationally expensive than that derived for
the Eulers angles. The price to pay is that 4 parameters are used, instead of only three,
that are not independent, inasmuch as they must satisfy the constraint
q02 + q q = 1
Moreover, their geometric interpretation during an evolution is less immediate than that
of the Eulers angles, the geometric meaning of which is intuitive. For this reason the
attitude of a satellite is often integrated in strapdown attitude determination systems in
terms of quaternions but then represented in terms of Euler angles.

1.2.6

Other attitude representations

The nonminimality of attitude representation in terms of quaternions can be solved for


by use of the Gibbs vector, defined as
tan(/2)
g = (g1 , g2 , g3 )T = a
Gibbs parameters, also known as Rodrigues parameters, are strictly related to quaternions,
as it is
(g1 , g2 , g3 )T = (q1 /q0 , q2 /q0 , q3 /q0 )T
The coordinate transformation matrix can also be represented in terms of Gibbs parameters:

1
LBI = (1 G)(1
+ G)
Being a minimal parametrization of attitudes, Gibbs parameters must present a singularity.3 The singular configuration of the Gibbs vector is for any eigenaxis rotation with
2

See above, pp. 326328


Recall that Euler demonstrated that any minimal parametrization of attitudes has at least one
singular configuration
3

G. Avanzini, Spaceflight Dynamics 1. Rigid Body dynamics

17

= , in which case their values diverge towards infinity. To solve for this problem, the
so called set of Modified Rodrigues parameters (MRP) was recently introduced, defined as
tan(/4)
p = (p1 , p2 , p3 )T = a
The singularity in the attitude representation is still present, and for a spinning body
the value of the MRPs will jump when crosses the critical threshol placed at half of
a rotation (that is, = ). Nonetheless the advantage of the MRPs is that they do
not diverge, but at the same time, no constraint on their value needs to be enforced in
order to save their meaning. These features make the numerical integration of MRPs less
critical with respect to both the quaternions case and the Gibbs vector.

1.3

Time derivative of vector quantities

If we consider a vector quantity in an inertially fixed reference frame FI ,


1 +YE
2 + ZE
3
~v = X E
its time derivative is given simply by
d~v
1 + Y E
2 + Z E
3
= X E
dt
that is

d~v
dt

n
oT
Y , Z
= X,
= v I

When the same vector quantity ~v is expressed in terms of components in a moving ref~ BI =
~ with respect to FI , the time
erence frame FB , rotating with angular velocity
derivatives of
~v = x
e1 + y
e2 + z
e3
is given by4
d~v
~ (x
= x
e1 +
e1 ) +
dt
~ (y
+ y
e2 +
e2 ) +
~ (z
+ z
e3 +
e3 )
This means that, in terms of vector components in FB , it is
 
dv
= v B + B v B
dt B
where
v B = {x,
y,
z}
T
B = {1 , 2 , 3 }T
4

Remeber the Poisson formula for the time derivative of a unit vector,
d
ei
~ e
i
=
dt

18

G. Avanzini, Spaceflight Dynamics 1. Rigid Body dynamics

3
E

3
e

~r
~r 0

dm

~v

2
e

CM

1
E

~
h

2
E
1
e

Figure 1.4: A rotating rigid body.

1.4
1.4.1

Eulers equations of motion of a rigid body


The inertia tensor

~ of a mass element m, moving with velocity ~v is


The angular momentum h
~ = ~r (m~v )
h
where ~r is the position vector of the mass, with respect of the pole used for the evaluation
of moments of vector quantities.
For an extended rigid body (Fig. 1.4), the total angular momentum is given by
Z
~ = (~r ~v )m
h
B

If the body is rotating around its center of mass, the velocity of every mass element is
~v =
~ ~r
so that
~ =
h

[~r (~
~r )] m

Expressing the vector quantities in body components as B = (1 , 2 , 3)T and r B =


~ ~r is given by
(x, y, z)T , the vector product
~ ~r = (2 z 3 y)

e1 + (3 x 1 z)
e2 + (1 y 2 x)
e3
Carrying on the calculations, the product ~r (~
~r) is


~r (~
1 +
~r ) =
(y 2 + z 2 )1 (xy)2 (xz)3 e


2
2
+
+ (xy)1 + (x + z )2 (yz)3 e

 2
2
2
3
+ (xz)1 (yz)2 + (x + y )2 e

G. Avanzini, Spaceflight Dynamics 1. Rigid Body dynamics

19

The integration over the body B of [~r (~


~r )] m is strictly function of the mass
distribution only, as angular velocity components are independent of body shape and
location. This means that, letting
~ = h1 e
1 + h2 e
2 + h3 e
3
h
it is
h1 = Ix 1 Ixy 2 Ixz 3
h2 = Ixy 1 + Iy 2 Iyz 3
h3 = Ixz 1 Iyz 2 + Iz 3
where the moments of inertia Ix , Iy , Iz , and the products of inertia Ixy , Ixz , Iyz , are
Z
Z
Z



2
2
2
2
Ix =
y + z m ; Iy =
x + z m ; Iz =
x2 + y 2 m
B
BZ
Z
ZB
Ixy =
(xy) m ;
Ixz =
(xz) m ;
Iyz =
(yz) m
B

In matrix form the angular momentum components in body axes are given by
hB = I B
where the symmetric matrix

Ix Ixy Ixz
I = Ixy Iy
Iyz
Ixz Iyz
Iz

is the inertia matrix that represent the inertia tensor in body axes.
The same relations can be derived directly in a more compact vector form remembering
that, for the double vector product, the following relation holds:
~ (~y ~z ) = (~
x
x ~z ) ~y (~
x ~y ) ~z
so that, in the present case, it is
~r (~
~ (~r
~ ) ~r
~r ) = (~r ~r )
Taking into account the definition of the dyadic tensor
(~
x~y ) ~z = (~y ~z )~
x
and the fact that the angular velocity vector is constant and can be taken out of the
integral, it is
Z
~
h =
[~r (~
~r )] m
B
Z

~
=
[(~r ~r ) 1 (~r~r )] m
B

~
= I

where I is the inertia tensor. Expressing ~r in terms of body components and integrating
over B the previous expression for the inertia matrix I is obtained.

20

G. Avanzini, Spaceflight Dynamics 1. Rigid Body dynamics

Principal axes of inertia


The matrix I is real and symmetric, so its eigenvalues are real5 and its eigenvectors are
mutually orthogonal.6 This means that there exists a body reference frame FP such that
the inertia matrix is diagonal,

Jx 0 0
I = 0 Jy 0
0 0 Jz

where the principal moment of inertia Jx , Jy , and Jz are the eigenvalues of I. The
eigenvectors are called principal axes.
Symmetries
If the mass distribution of the body B is characterized by symmetries, this property
reflects onto the inertia matrix I. As an example, if B has a plane of symmetry, one of the
principal axis will be perpendicular to the plane and the other two will lie on that plane.
If this case, the products of inertia relative to the axis perpendicular to the symmetry
plane will be zero. This case is typical of fixed wing aircraft, where the longitudinal
plane is approximately a symmetry plane. The y body axis, directed perpendicular to the
symmetry plane, is characterized by zero products of inertia, so that the inertia matrix
5

From the definition of eigenvalue and eigenvector of a complex matrix A, it is easy to derive the
following equation,
xH Ax
Ax = x = H
x x
If A is Hermitian (i.e. the linear operator represented by A is selfadjoint), it is
y H Ax = (Ay)H x
so that, remembering the properties of the hermitian inner product, such that xH y = (y H x), it is
H
H
H
= x Ax = (Ax) x = x Ax =

H
H
x x
x x
xH x

which means that the eigenvalue is equal to its conjugate, i.e. it must be real.
6
Given two distinct eigenvalues i 6= j and their respective eigenvectors xi and xj , the following
relations hold:
Axi

= i xi

Axj

= j xj

H
Multiplying the first equation by xH
j and the second one by xi , taking the complex conjugate of the
second and subtracting it from the first, one gets
T
H
T
xH
j Axi (xi Axj ) = i xj xi (j xi xj )

Remembering that the eigenvalues are real, it is


H
H
xH
j Axi (Axj ) xi = (i j )xj xi

Taking into account the definition of Hermitian operator the first term of the last equation is zero and so
the Hermitian product xH
j xi is zero if i 6= j . Since both xj and xi are real, their Hermitian product
coincides with the scalar product, so that distinct eigenvectors are real and perpendicular.

G. Avanzini, Spaceflight Dynamics 1. Rigid Body dynamics

21

of an aircraft is typically equal to

Ix
0 Ixz
Iy
0
I= 0
Ixz 0
Iz
If the body is axi-symmetric (or simply has a regular polygonal mass distribution
w.r.t. an axis ), the symmetry axis is a principal axis of inertia, while any couple of
perpendicular axes on the plane normal to will complete the set of principal axes.
In this case the principal moments of inertia relative to the axes perpendicular to the
3 is a symmetry axis, the inertia
symmetry axis will be equal. Assuming that = e
tensor becomes

Jt 0 0
I = 0 Jt 0
0 0 Js
where the subscripts t and s indicate the transverse and spin (or axial) moments of
inertia, respectively.

1.4.2

Rotational kinetic energy

The kinetic energy of a body is given by


Z

1
T =
2

(~v ~v ) m

~ ~r , the argument of the integral


Remembering that for a rotating rigid body it is ~v =
becomes (~
~r ) (~
~r). Also, taking into account the permutation property of the
scalar triple product
~ (~y ~z ) = ~y (~z x
~ ) = ~z (~
x
x ~y )
one obtains the equivalence
~ [~r (~
(~
~r ) (~
~r ) =
~r )]
Substituting this expression into the integral, and taking the (constant) angular velocity
out of the integration symbol, one gets
1
~
T =
2

[~r (~
~r )] m

that, remembering the definition of the inertia tensor, brings


  1
1
~ I
~ = TB (I B )
T =
2
2
or, equivalently,
1
~ = 1 T hB
~ h
T =
2
2 B

22

1.4.3

G. Avanzini, Spaceflight Dynamics 1. Rigid Body dynamics

Eulers equation of motion

The second fundamental law of rigid body dynamics states that the time derivative of
the angular momentum is equal to the external torque applied to the body B. In vector
form, it is
~
dh
~
=M
dt
Expressing the vector quantities in body axis components one gets
h B + B hB = M B
If the inertia matrix I is constant, it is
I B + B (I B ) = M B
When a set of principal axes is chosen as the body axes, the inertia tensor is diagonal
and the Eulers equation of motion for a rigid body are obtained:
Jx 1 + (Jz Jy )2 3 = M1
Jy 2 + (Jx Jz )3 1 = M2
Jz 3 + (Jy Jx )1 2 = M3
These equations can be integrated as a function of the applied torque to obtain the
time history of the angular velocity components. These, in turn, can be used to determine
the variation with time of the Eulers angles (or of the quanternions), thus describing the
evolution of the rigid body attitude.

1.4.4

Conservation of angular momentum

Writing Eulers equations in a set of principal axes such that (without loss of generality)
Jx > Jy > Jz , torquefree motion is described by the following set of ODEs,
Jx 1 + (Jz Jy )2 3 = 0
Jy 2 + (Jx Jz )3 1 = 0
Jz 3 + (Jy Jx )1 2 = 0
It is easy to demonstrate that the magnitude h of the angular momentum vector
~ = h1 e
1 + h2 e
2 + h3 e
3
h
1 + Jy 2 e
2 + Jz 3 e
3
= Jx 1 e
is constant when a torquefree motion is considered. A first intuitive derivation of this
property is that if the applied torque vanishes the angular momentum vector is constant in
FI , and its magnitude is independent of the considered reference system. It is also possible
to demonstrate analytically that h = ||h|| is constant, by taking the time derivatives of
h2 = (Jx 1 )2 + (Jy 2 )2 + (Jz 3 )2
in the hypothesis of torquefree motion (M1 = M2 = M3 = 0),

dh2
= 2 Jx2 1 1 + Jy2 2 2 + Jz2 3 3
dt
= 2 [Jx 1 (Jy Jz )2 3 + Jy 2 (Jz Jx )1 3 + Jz 3 (Jx Jy )1 2 ]
= 2 (Jx Jy Jx Jz + Jy Jz Jy Jx + Jz Jx Jz Jy ) 1 2 3 = 0

G. Avanzini, Spaceflight Dynamics 1. Rigid Body dynamics

23

Geometrically, the angular velocity vector must lie on an ellipsoid (the angular momentum
ellipsoid ) in FB , the equation of which takes the form
12
22
32
+
+
=1
(h/Jx )2 (h/Jy )2 (h/Jz )2

1.4.5

Conservation of kinetic energy

In a similar way, it is also possible to demonstrate that the kinetic energy of a rigid body
is constant if no external torque is applied. Again, taking the time derivative of

1
1
T = B h =
Jx 12 + Jy 22 + Jz 32
2
2

it is
dT
dt

= Jx 1 1 + Jy 2 2 + Jz 3 3
= 1 (Jy Jz )2 3 + 2 (Jz Jx )1 3 + 3 (Jx Jy )1 2
= 2 (Jy Jz + Jz Jx + Jx Jy ) 1 2 3 = 0

This means that the angular velocity vector must satisfy also the equation
22
32
12
+
+
=1
(2T /Jx ) (2T /Jy ) (2T /Jz )
that is, it must point the surface of the kinetic energy (or Poinsot) ellipsoid in FB . The
combination of these two last results demonstrate that the angular velocity vector describe
the curve given by the intersection of the angular momentum ellipsoid and the kinetic
energy ellipsoid, which is called the polhode.

1.5

Generalised Euler equation

In their original formulation, Euler equations are written in a bodyfixed reference frame
FB with the origin in the center of mass O of the body B. On one side, the expression
employed for the angular momentum h of B requires that B is rigid, and centering B in
O greatly simplify the expression. At the same time, linear and angular momentum conservation laws do apply to any mechanical system (under sufficiently mild assumptions),
so that it is possible to obtain a generalised formulation for the equation of motion in a
frame which is non centered in O.

1.5.1

Derivation of the generalised form of Euler equation

The classical equations, referred to the center of mass O, are


m~aO = F~
~O
dh
~O
= M
dt
where, m is the mass of B and F~ is the external force, producing an acceleration ~aO of
the center of mass. Considering an arbitrary point A with arbitrary motion with respect

24

G. Avanzini, Spaceflight Dynamics 1. Rigid Body dynamics

to the body, such that ~r AO is the position vector of the centre of mass O with respect to
A, the torque acting on the body can be referred to A,
~ A=M
~ O + ~r AO F
~
M
As for the angular momentum relative to A, it is

Z 
d~r AP
~
~r AP
hA =
m
dt
B
where ~rAP is the position vector of the mass element m with respect to A, while d~r AP /dt
its (relative) velocity. Upon substitution of ~r AP = ~r AO + ~r OP into the expression for ~hA ,
one obains

Z 
d
~
hA =
(~r AO + ~r OP ) (~r AO + ~r OP ) m
dt
B


Z 
Z 
d~r AO
d~r OP
~r AO
~r AO
=
m +
m +
dt
dt
B
B


Z 
Z 
d~r OP
d~r AO
~r OP
~r OP
m +
m
+
dt
dt
B
B
The last term is the angular momentum with respect to the centre of mass

Z 
d~r OP
~
~r OP
hO =
m
dt
B
while the second and the third one are both zero, from the definition of centre of mass.7
~ A as
It is thus possible to write h
~A = h
~ O + m~r AO d~r AO
h
dt
The acceleration of O can be rewritten as a function of the absolute acceleration of A
~aO = ~aA +

d2~r A
dt2

By substituting the above expressions in the angular momentum equation, one gets


 

2
~
d ~
d~r A
d
r
A
~ A ~r A m ~aA +
hA m~r A
=M
dt
dt
dt2
which can be rewritten as

~A
dh
~ A ~aA = M
~ A
+S
dt
~ A = m~r A is the static moment of the body with respect to A.
where S
7

If m is the total mass of the body, the absolute position of the centre of mass is defined as
Z
1
~r O =
~rOP m
m B

so that, letting the position vector of P with respect to O be defined as ~r OP = ~r P ~r O , it is


Z
~r OP m = 0.
B

Taking into account that integration is a linear operator, that can be commuted with the time derivative,
both the second and the third terms vanishes.

G. Avanzini, Spaceflight Dynamics 1. Rigid Body dynamics

1.5.2

25

Use of the generalised form of Euler equation

The generalised form of classical Euler equation allows for writing dynamic models of
systems where the mass distribution is not constant. On one side, the assumption of
rigidity often applies with reasonable accuracy to many spacecraft, at least over a relatively
long timescale. As an example, the slow consumption of fuel during the whole operational
lifetime of the vehicle (years) causes a shift of the centre of mass, but the knowledge of
the current value is usually sufficient for describing manoeuvers over a faster timescale
(minutes).
On the converse, it is sometimes necessary to take into account phenomena where
variations of the mass distribution affects significantly the vehicles attitude dynamics.
Flexible solar panels or other appendages or fuel sloshing in the tanks, the motion of
which can be excited by an attitude or an orbit manoeuvre, may cause as a reaction
oscillations of the spacecraft with respect to the desired attitude, which may harm overall
pointing accuracy, unless properly damped passively by the inherent vehicle stability or
actively by its automatic control system. Sometimes, a satellite may feature a nutation
damper, where a mass moving within a viscous liquid induces dissipation in order to
asymptotically stabilise a pure spin condition (see Chapter 3).
If the centre of mass O does not maintain a fixed position with respect to the body
and one must choose whether (i) keeping the reference frame centred in O or (ii) choosing
another reference point, allowing for displacements of O with respect to the origin of the
frame. In the second case, the generalised form of Euler equation can be employed for
the angular momentum balance, the current position of the centre of mass and the inertia
tensor being usually available from the knowledge of the (current) mass distribution.
The major advantage of the second approach lies in the fact that it is often possible to
identify a frame which is fixed with respect to the rigid structure of the spacecraft, which
allows to describe intuitively the orientation of the spacecraft itself and its equipment
(antennae, sensors, and so on). If the first approach is taken, the pseudobody frame
moves with respect to the spacecraft structure because of changes in the mass distribution,
so that the knowledge of its attitude does not imply the knowledge of spacecraft and
sensor orientation. Moreover, the expressions of the attitude equations are usually rather
complex.

26

G. Avanzini, Spaceflight Dynamics 1. Rigid Body dynamics

Chapter 2
Dynamics of Spinning Spacecraft
Spin stabilisation in a simple, low cost and effective means of attitude stabilisation. Prior
to, or just after deployment, the satellite in spun up about its axis of symmetry. For
this reason, spin stabilised satellites are usually short cylinders. The angular momentum accumulated about the spin axis the provides gyroscopic stability against external
disturbance torques.
Although simple and reliable, spin stabilised satellites are inefficient for power generation. Since the satellite is continually spinning, the entire surface of the satellite must
be covered with solar cells. In addition, payload efficiency is particularly low when only
one direction is fixed in space and maneuvering of the spin axis complex.
Most of these drawbacks were solved by the introduction of a dualspin satellites,
where a part of the satellite (the socalled platform) is despun and brought at rest with
respect to the desired reference frame, while gyroscopic stability to the whole system is
provided by the spinning rotor. Payload efficiency is greatly enhanced as it is usually
possible to pivot antennae or sensors with respect to the (fixed) platform, thus gretly
reducing the need for manoeuvres, unless it is necssary to aim them towards a target in
a completely different direction.

2.1

Torquefree motion of axisymmetric satellites

The principal moments of inertia of an axisymmetric satellite will be given by


Jt = Jx = Jy
Js = Jz
where the subscripts t and s stand for transverse and spin, respectively, and we assume
that the symmetry axis coincides with the third (
e3 ) axis of the body frame FB .
During the spinup manoeuver, the satellite will accumulate angular momentum about
the spin axis, but because of various perturbations or imperfections, such as thruster
misalignment, the final condition at the end of the spinup will hardly be a pure spin
3 . The imperfections will cause some (hopefully residual) nutation.
about the spin axis e
For torquefree motions
M1 = M2 = M3 = 0
of axial symmetric spacecraft, Eulers equations take the following form:
Jt 1 + (Js Jt )2 3 = 0
Jt 2 + (Jt Js )1 3 = 0
Js 3
= 0
27

28

G. Avanzini, Spaceflight Dynamics 2. Dynamics of Spinning Spacecraft

The first two equations are coupled, while the third one is independent of the other
two. This means that the latter one can be integrated on its own. The resulting (trivial)
solution is given by
3 =
where is the (constant) spin rate about the spin axis.
Letting
Js Jt
=

Jt
the first two equations can be rewritten as
1 + 2 = 0
2 1 = 0
Multiplying the first equation by 1 and the second by 2 and summnig up, one gets
1 1 + 2 2 = 0
that is
2
12 + 22 = 12
= constant

~ that lies in the e


1 e
2
This means that the component of the angular velocity vector
plane, namely
~ 12 = 1 e
1 + 2 e
2

has a constant magnitude. As also 3 is constant, we get that


2
||~
|| = 12 + 22 + 32 = 12
+ 2 = constant

The first two equations of motion,


1 + 2 = 0
2 1 = 0
can be easily integrated. In fact, deriving the first one with respect to the time t, one gets

1 + 2 = 0
that, taking into account the second equation, becomes

1 + 2 1 = 0
which is formally identical to the well known equation of the linear harmonic oscillator.
The general solution
1 (t) = A cos(t) + B sin(t)
for initial conditions
1 (t = 0) = 1,0 ; 1 (t = 0) = 1,0
becomes
1 (0)
sin(t)

= 12 sin[(t t0 )]

1 (t) = 1 (0) cos(t) +

G. Avanzini, Spaceflight Dynamics 2. Dynamics of Spinning Spacecraft

29

[rad s1]

0.1
0.05
0
0.05
0.1

[rad s1]

0.1
0.05
0
0.05
0.1

3 [rad s ]

0.15
0.1
0.05
0

0.5

t [103 s]

Figure 2.1: Timehistory of angular velocity components for a free spin condition.
Deriving the solution for 1 w.r.t. t and substituting into the first equation, it is
2 =

1
1 (0)
= 1 (0) sin(t)
cos(t)

= 12 cos[(t t0 )]

3 with angular velocity .


The evolution of 1 and 2 shows that ~12 whirls around e
Writing the angular velocity as
~ =
~ 12 +

e3
~ describes a cone around the axis of symmetry e3 of the spinning
during the evolution,
body, which is called the body cone.
The angular momentum vector can be written in the form
~ = J1 1 e
1 + J2 2 e
2 + J3 3 e
3
h
1 + 2 e
2 ) + Js 3 e
3
= Jt (1 e
~ 12 + Js
= Jt
e3
~ and
~ are both a linear combination of the vectors
~ 12 and
It can be observed that h
~
3 . Thus, during the motion of the spinning body, the vectors h,
~ and e
3 lie in the same
e
3 , if we look at the motion from FB .
plane , that rotates around e
~ and
~ are not aligned. They are aligned only if 12 = 0,
In the most general case h
that is, if we have a pure spinning motion about the symmetry axis. This is the desired
spin condition, where a sensor placed on the satellite on its symmetry axis points a fixed
direction in space. If 12 6= 0, the motion of the spinning body is more complex (and the
pointing less accurate). A geometric description of this motion will now be derived.
~ = 0, implies that
The assumption of torquefree motion, M
~
dh
~ = constant
~ = 0 = h
=M
dt

30

G. Avanzini, Spaceflight Dynamics 2. Dynamics of Spinning Spacecraft


AA
~ A
h

)

B
B

PP
P
q
P
PP
q

3
e

~ 12
h
~ 12

1
e

2
e

Figure 2.2: Torquefree spinning of an axisymmetric satellite as seen in FB .


~ if we look at the motion from FI .
in the inertial frame, and rotates around h,
It is possible to define two angles, and , that remains constant in time,

h12
Jt 12

tan =
=

Jt
h3
Js 3
= tan =
tan

12
Js

tan =
3

~ is the (constant) nutation angle and


3 k h,
If we choose an inertial frame such that E
3 in the inertial space. The angle is the
it defines the orientation of the symmetry axis e
~ and
~ , equal to
semiaperture of the bodycone. This means that the angle between h
~ fixed in the inertial frame, the
~ describes a cone around h,
| |, is also constant, and
space cone.
The body cone is attached to the body axes, but it is not fixed in space. On the
~ that is constant in the inertial frame,
converse, the space cone, attached to the vector h,
~ , that is the axis of instantaneous
is fixed in FI . The two cones remains tangent along
rotation of the body, and the motion of the satellite can be represented by the body cone
rolling along the surface of the space cone.
As a final observation, it should be noted that, when Js > Jt (oblate body, that is
a diskshaped body), it is > , and the space cone is inside the body cone. On the
converse, when Js < Jt (prolate body, that is a rodshaped body), it is < , and the
space cone is outside the body cone.
For what concern the attitude resulting from such a motion, substituting the expres-

G. Avanzini, Spaceflight Dynamics 2. Dynamics of Spinning Spacecraft

31

3 k h
~
E
~

3
e

1
E

2
E

Figure 2.3: Torquefree spinning of an axisymmetric satellite as seen in FI .


sions for the angular velocity components determined above in the following equation,
sin sin +
cos
1 =
cos sin
sin
2 =
cos +
3 =
= 0, one gets
and remembering that = constant
sin sin = 12 sin[(t t0 )]

cos sin = 12 cos[(t t0 )]

cos + =

By squaring and summing the first two equations, it is evident that


2 sin2 = (12 )2

is constant;
is called precession rate or coning speed, and it is the angular
so that
velocity of the line of the nodes on the horizontal plane.
Dividing the first equation by the second, the spin angle is determined,
tan = tan[(t t0 )] = (t t0 )
Deriving w.r.t. time, the inertial spin rate is obtained
=
that is also constant. At this point, it is possible to evaluate the precession rate from the
third equation,

cos + =
=

cos

32

G. Avanzini, Spaceflight Dynamics 2. Dynamics of Spinning Spacecraft

But from the definition of ,


=

Js Jt
Jt
Jt
=
=

Jt
Js Jt
Jt Js

one gets
=

Js

Js
=
Jt Js cos
Jt cos

and have the same sign and we have


If Jt > Js , that is we have a prolate body,
direct precession, that is the precession rate is in the same direction of the spin rate. On
and have different signs
the converse, if Js > Jt and an oblate body is dealt with,
and we have retrograde precession, the precession rate being in the opposite direction with
respect to the spin rate.

An observation
It is important to note that the derivation presented in this paragraph are valid for any
rigid body which has two equal principal moment of inertia. This is a category much wider
than that of axisymmetric bodies, including any prism with a basis made of a regular
polygon, but also any other body of irregular shape such that there exists a set of principal
axes of inertia such that Jx = Jy 6= Jz .
The purespin and the flatspin condition
When the nutation angle is zero, the symmetry axis, the angular velocity and the angular
momentum vectors are all parallel. This situation, which is referred to as the purespin
condition, is an equilibrium for the system, as it can be easily demonstrated by checking
that for B = (0, 0, )T the derivatives of all the components of the angular velocity
vector are zero. This means that in the purespin condition the symmetry axis is (and
remains!) fixed in the inertial space, as far as no transverse angular velocity component is
present, unless an external perturbation affects the satellite dynamics. As a consequence,
this is the desired condition for the axisymmetric spinning spacecraft. Any displacement
~ would cause a precession rate, that is, the symmetry axis
3 from the direction of h
of e
~ and the sensors mounted along e
3 will
rotates about the (inertially fixed) direction of h
not point towards a fixed direction in space.
The purespin condition is stable, in the sense of Lyapunov: A small displacement
from the ideal case will induce precession, but the nutation angle remains constant. This
means that no divergence takes over and the pointing error remains finite and, possibly,
small, for both prolate and oblate bodies. Such an ideal situation will be dramatically
influenced by the effects of energy dissipation, as described in Section 2.3.
It can be seen the also B = (1 , 2 , 0)T is an equilibrium. In this case the satellite
3 will induce
spins about a transverse axis. As far as any minor perturbation about e
a constant (although slow) drifting motion about the symmetry axis, this condition is
unstable in the sense of Lyapunov, inasmuch as 1 and 2 will vary as the sine and
the cosine of t, thus leaving the neighbourhood of the equilibrium. Also in this case,
dissipation will take its toll.

G. Avanzini, Spaceflight Dynamics 2. Dynamics of Spinning Spacecraft

33

~
h

d
O

Figure 2.4: The invariable plane.

2.2

Torquefree motion of triinertial satellites

An analytical solution of the motion of triinertial rigid body can be derived in terms of
Jacobi elliptic functions. Luckily there is also a geometric description of the same motion,
due to Poinsot, which is much simpler, nonetheless extremely useful for the description
of the dynamics of an arbitrary rigid body.
Remembering that the kinetic energy
1
~
~ h
T =
2
~ are constant, it is possible to consider the (constant)
and the angular momentum vector h
dot product
~
h
2T
~ =

=d
(2.1)
h
h
~ (Fig. 2.4). The
as the length d of a (constant) segment ON along the direction of h
~ placed at a
invariable plane , which is the plane perpendicular to the direction of h,
distance d from the body center of mass O, is thus fixed in FI , and it represents the locus
~ that satisfy Eq. (2.1). Remembering that the Poinsot ellipsoid is
of all the possible
~ that satisfy the kinetic energy equation, the intersection
the locus of all the possible
between the ellipsoid and the invariable plane must contain the angular velocity vector.
It is easy to demonstrate that such an intersection is a single point P , i.e. the Poinsot
ellipsoid and the invariable plane are tangent. Since the time derivative of the kinetic
energy is zero,
1 d~
~
dT
=
h=0
dt
2 dt
~ are perpendicular, thus d~
~ + d~
the increment d~
and h
lie on . But since the vector

must be also on the Poinsot ellipsoid, d~


must be tangent to its surface. These two
conditions can be satisfied only if the Poinsot ellipsoid is always tangent to . Moreover,
the Poinsot ellipsoid is fixed in the body frame FB , so that d~
is the same in both FI
and FB . This means that the Poinsot ellipsoid rolls without slipping on , as depicted in
Fig. 2.5.

34

G. Avanzini, Spaceflight Dynamics 2. Dynamics of Spinning Spacecraft

~
h


+
d~

~

d
O

Figure 2.5: The Poinsot ellipsoid rolls on the invariable plane.


~ s on the Poinsot ellipsoid is given by the polhode
The locus of all the possible
curve, which is the intersection between the Poinsot ellipsoid and the angular momentum
ellipsoid. Thus, during the rolling motion, the tangent point moves on the Poinsot ellipsoid
along the polhode. The corresponding curve on is the herpolhode. When the body is
axisymmetric, both the polhode and the herpolhode are circles and the situation can be
depicted in terms of space and body cones. In general the herpolhode is not a closed
curve, but the polhode must be a closed curve on the Poinsot ellipsoid, inasmuch as after
~ must attain again the same value, in order
a revolution around the spin axis, the vector
to satisfy conservation of both kinetic energy and angular momentum.

2.2.1

Drawing the polhode curves

In order to draw the shape of the polhodes on the Poinsot ellipsoid, it is sufficient to recall
the equations of the Poinsot ellipsoid and the angular momentum ellipsoid, that are
12
22
32
+
+
= 1
(2T /Jx ) (2T /Jy ) (2T /Jz )
12
22
32
+
+
= 1
(h/Jx )2 (h/Jy )2 (h/Jz )2
Subtracting the first equation from the second and multiplying the result by h2 , one gets






h2
h2
h2
2
2
Jx Jx
1 + Jy Jy
2 + Jz Jz
32 = 0
2T
2T
2T
the polhode equation. In order to have real solutions for the above equation, the three
coefficients cannot have the same sign. For this reason the parameter J = h2 /(2T ) must
lie between the maximum and the minimum moment of inertia. Assuming, without loss
of generality, that Jx > Jy > Jz , it is
Jz J Jx

G. Avanzini, Spaceflight Dynamics 2. Dynamics of Spinning Spacecraft

35

2
1
0
1
2
2
1

1
2

1
2

Figure 2.6: Polhode curves on the Poinsot Ellipsoid.


The easiest way to determine the shape of the polhodes is to consider their projection
onto the planes of the threedimensional space 1 2 3 . Eliminating 3 between the
equations of the two ellipsoids brings the equation
Jx (Jx Jz )12 + Jy (Jy Jz )22 = 2T (J Jz )
which represents an ellipse, since all the coefficients are positive. Similarly, eliminating
1 one gets
Jy (Jy Jx )22 + Jz (Jz Jx )32 = 2T (J Jx )
which is again the equation of an ellipse, inasmuch as all the coefficients are negative. On
the converse, eliminating 2 , which is the angular velocity component with respect to the
intermediate axis, brings
Jx (Jx Jy )12 + Jz (Jz Jy )32 = 2T (J Jy )
which represent a hyperbola, the coefficients of the lefthand side being of different sign.
It should be noted that, depending on the sign of J Jy , which can be either positive
or negative, the axis of the hyperbola will be vertical or horizontal, in the 1 3 plane.
When J = Jy , the polhode equation degenerates into the form
Jx (Jx J ) 12 + Jz (Jz J ) 32 = 0
which represents the separatrices, the boundaries of motion about the axis of maximum
and minimum inertia.

36

2.2.2

G. Avanzini, Spaceflight Dynamics 2. Dynamics of Spinning Spacecraft

Stability of torquefree motion about principal axes

Spinning about any of the principal axis is an equilibrium condition for a rigid body of
arbitrary inertias. In this case, the angular momentum vector, the angular velocity vector
and the principal axis used as the rotation axis are all parallel and the time derivatives
of all the components of B are zero. The three spin condition x = (, 0, 0)T , y =
(0, , 0)T , and z = (0, 0, )T are the purespin condition for the threeinertial body.
The shape of the polhodes already provide an information about the stability of these
equilibria. The axes of maximum and minimum inertia are centres surrounded by finite
size closed orbits, which means that a perturbation will induce a motion in the neighbourhood of the original equilibrium. On the converse, the intermediate axis of inertia is a
saddle point, that is a small perturbation will take the angular velocity vector far from
the initial point in the neighborhood of the saddle, so that spinning around the intermediate axis is an unstable condition for the satellite, and cannot be used for gyroscopic
stabilisation.
These facts can be demonstrated analytically. Let us consider a spinning motion about
the z axis of the principal frame (this time without making assumptions on the order of
~ 0 =
~ =
~ 0 + ~
the principal moment of inertia), such that
e3 . Assuming that
,
T
where the body frame components of ~
, given by B = {1 , 2 , 3 } , are small
perturbations with respect to , Eulers equations can be rewritten as follows
Jx 1 + (Jz Jy )2 = 0
Jy 2 + (Jx Jz )1 = 0
Jz 3 = 0
where second and higher order terms were neglected.
The third equation is decoupled, thus stating that a perturbation on the spinning axis
does not affect the other two. The first two equations can be rewritten in matrix form

 


1
0
(Jy Jz )/Jx
1
=
2
(Jz Jx )/Jy
0
2
The characteristic equation is thus given by
2 + 2

(Jy Jz )(Jx Jz )
=0
Jx Jy

and its roots are pure imaginary,


s

= i

(Jy Jz )(Jx Jz )
Jx Jy

if either Jy < Jz and Jx < Jz , or Jy > Jz and Jx > Jz , that is if the spin axis is either
the axis of maximum or minimum moment of inertia. In such a case, the solution of the
perturbation equation is confined in the neighborhood of the spin condition and divergence
because of neglected nonlinear terms is ruled out on energy conservation grounds, so that
these conditions are Ljapunov (although not asymptotically) stable. On the converse, if
the spin axis is the intermediate one, the product (Jy Jz )(Jx Jz ) is negative, inasmuch
as one of the two factor is positive and the other is negative. In this case, the roots are
both real,
s
(Jy Jz )(Jx Jz )
=
Jx Jy

G. Avanzini, Spaceflight Dynamics 2. Dynamics of Spinning Spacecraft

37

and one of the two eigenvalues is strictly positive. This means that spinning around the
intermediate axis is an unstable equilibrium for the spinning rigid body.

2.3

Nutation damping

Nutation damping is a simple yet effective way to restore a state of pure spin, if a nutation
angle different from zero should be induced by some external cause. We will now investigate how and under which circumstances it is possible to exploit the effects of energy
dissipation in order to make a pure spin condition asymptotically stable.

2.3.1

Effects of energy dissipation on axisymmetric satellites

As a matter of fact, the rigid body is an abstraction. Usually flexible appendages and/or
fuel sloshing induce some energy dissipation the effects of which can be easily determined
at least in the case of axisymmetric satellites by use of the energy sink model.
The kinetic energy of a rotating rigid body is given by
1
T = (Jx 12 + Jy 22 + Js 32
2
The expression for T can be rearranged for the axisymmetric case in the form
1
1
T = Jt (12 + 22 ) + Js 2
{z
} |2 {z }
|2
transverse
spin

At the same time, the angular momentum is

~ = Jt 1 e
1 + Jt 2 e
2 + Js 3 e
3 h2 = Jt2 (12 + 22 ) + Js2 2
h
so that the quantity (2Js T h2 ) is
(2Js T h2 ) = Js Jt (12 + 22 ) + Js2 2 Jt2 (12 + 22 ) Js2 2
= Jt (12 + 22 )(Js Jt )
By taking the time derivatives of both sides of the equation, and remembering that,
for zero external torque the angular momentum vector remains constant, one gets
dT
1 Jt
d
=
(Js Jt ) (12 + 22)
dt
2 Js
dt
If there is no dissipation, T = 0 and 12 + 22 is constant (as already demonstrated in
Section 2.1). If there is dissipation, T < 0, which means that, assuming an approximately
constant mass distribution, so that the moments of inertia are unaffected, 12 + 22 must
either grow or decrease. It should be noted that, if the body is oblate, than Js Jt > 0 and
d(12 + 22 )/dt < 0, which meand that the transverse component of the angular velocity
vector is decreasing. Dissipation will stop only when it is 12 + 22 = 0, that is, when a
pure spin condition is reached. In such a case, dissipation turns the pure spin condition
into an asymptotically stable equilibrium: a displacement from such a condition will be
damped out at the expenses of a reduction of the system energy.
On the converse, if the body is prolate and Js Jt < 0, than dissipation induces an
increase of the transverse angular velocity component, as in order to satisfy the equation,

38

G. Avanzini, Spaceflight Dynamics 2. Dynamics of Spinning Spacecraft

[rad s1]

0.1
0.05
0
0.05
0.1

[rad s1]

0.1
0.05
0
0.05
0.1

3 [rad s ]

0.15
0.1
0.05
0

0.5

t [103 s]

Figure 2.7: Timehistory of angular velocity components with nutation damping.


d(12 + 22 )/dt must be positive. This means that the nutation angle is growing, being
proportional to 12 /. In such a case the pure spin condition about the symmetry axis,
which was stable in the sense of Lyapunov for the perfect rigid case, becomes unstable.
Although divergence is usually slow, the nutation angle will steadily increase with time.
The rate of variation of the nutation angle is slow enough to make its rate negligible
Under this assumption it is possible to state that
with respect to .
2 sin2
12 + 22
and, as a consequence, it is possible to formulate the energy decay rate as
dT
1 Jt
d 2 2
=
(Js Jt ) (
sin )
dt
2 Js
dt
Moreover, for small nutation angles the precession rate can also be considered approximately constant, and differentiation with respect to time bring the following expression:
Jt 2
dT

=
(Js Jt ) sin cos
dt
Js
must also be
When energy is dissipated and T < 0, the quantity (Js Jt ) sin(2)
negative, and a nutation rate is induced by dissipation. As qualitatively shown previously,
is negative for small nutation angles, and
if the satellite is an oblate body and Js > Jt ,
tends to decrease with time because of dissipation. This means that spinning around the
symmetry axis is asymptotically stable, as 0. For a prolate body, when Js < Jt , the
resulting nutation rate is positive and the nutation angle increases with time: the satellite
tends towards a flat spin condition, /2.
Note that both the purespin and flatspin conditions are always equilibria for the
= 0 if either = 0 or /2,
satellite, regardless of its shape, inasmuch as T = 0 and
but the stability of these equilibria is affected by the mass distribution.

G. Avanzini, Spaceflight Dynamics 2. Dynamics of Spinning Spacecraft

39

3
e

2
e

(md /m)
O


+
u

~b

1
e

md

~
n
Figure 2.8: A satellite equipped with a ballintube damper.
Also note that as 1 and 2 are damped out in the stable case, 3 increases because
of conservation of angular momentum. Letting f be the final spin rate, it is
h2 = Jt2 (12 + 22) + Js2 2 = Js2 2f
Therefore,
2f

Jt2 2
= + 2 (1 + 22 )
Js
2

and the final spin rate results higher than the initial one, as expected (Fig. 2.7). In the
unstable case, dissipation will make the spin angular velocity component drop towards
zero as the angular momentum is transferred by dissipation in the transverse direction.

2.3.2

More accurate models

In a real system there are several causes that induce energy dissipation, such as fuel
sloshing. Most satellites use thruster to actively remove nutation and maintain the desired
pure spin condition, when the nutation angles increases beyond a prescribed threshold.
For oblate bodies a nutation damper provides a very simple and useful means for passively
stabilizing the pure spin motion. An example of damper is the ballintube: a plastic tube
is filled with viscous fluid and a ball bearing is free to move in the fluid, thus dissipating
energy. The only drawback of this technique is that it may take several minutes (if not
hours) to remove nutation.
In order to derive the equations of motion of a satellite equipped with a springmass
dashpot damper, it is necessary to consider the generalized form of the Euler equation,
~
dh
~ ~aA = M
~
+S
dt

40

G. Avanzini, Spaceflight Dynamics 2. Dynamics of Spinning Spacecraft

written in a reference frame FA attached to the satellite platform P, with origin in the
point A, which is the center of mass of the satellite when the damper mass md is in the
undeformed position P ( = 0) for the spring (Fig. FigSatDamp). All the moments are
taken with respect to A, while moment of inertia and position of the centre of mass O are
affected by mass displacement along the tube.
The derivation of the complete set of equations of motion of the satellite equipped with
a damper is out of the scope of the course, but the results obtained from the analysis of the
model are of paramount importance. For a threeinertial satellite one finds out that, in
presence of energy dissipation, the only stable purespin condition takes place about the
axis of maximum inertia. Dissipation makes such a spin condition asymptotically stable,
that is, it transforms the closed curves around the reference spin condition into converging
spirals. On the converse, spin about the axis of minimum inertia is transformed into an
unstable equilibrium, the closed curves surrounding it being transformed into diverging
spirals. The instability of the spin condition about the intermediate axis is unaffected
by dissipation. The spacecraft will simply converge towards a spinning motion about the
axis of maximum inertia, unless some external action prevents it from doing so.

2.4

Attitude Maneuvers of a spinning satellite

As stated since the beginning of the chapter, one of the major drawbacks of spin stabilisation lies in the complexity of the technique necessary for performing attitude manoeuvres,
that is, changing the direction of the spin axis of the satellite in a controlled way. In order
to rotate the spin axis of a spinning spacecraft, it is possible to exploit the precession
rate induced by a properly sized nutation angle. This requires a variation of the angular momentum vector that can be obtained by applying a proper control torque to the
spacecraft.
~ at position ~r in the body frame centered
If we consider a thruster generating a force F
in the center of mass O, the resulting moment acting on the satellite is
~ = ~r F
~
M
Given a firing time of ton seconds, starting at t0 , the overall change of angular momentum
is equal to
Z
t0 +ton

~ =
h

t0

~ )dt.
(~r F

If the firing time is sufficiently small with respect to the period of rotation of the spinning
~ as
spacecraft Tspin = 2/Omega, it is possible to approximate h
~ = (~r F
~ )ton
h

The control torque is usually delivered by use of coupled pairs of thrusters, in such
a way that a net torque is produced, with no force perturbing the orbit motion of the
spacecraft.
The torque pulse changes almost instantaneously the direction of the spacecrafts
angular momentum vector, thus creating a nutation angle . At this point the spacecraft
is no longer spinning around its symmetry axis, and a precession rate builds up, equal to
=

Js

Jt Js cos

G. Avanzini, Spaceflight Dynamics 2. Dynamics of Spinning Spacecraft

h1 h
2

h1

h1

h1
h0

h0


:

h2


: X

z
X



+



+

F
O

41

F
O

Figure 2.9: Changing the spin axis direction with nutation angle and precession rate.
It must be remembered the the direction of the precession motion depends on whether
an oblate or a prolate body is considered, but in the latter case the instability of the spin
condition around the axis of minimum inertia when dissipation is present prevents the
use of this technique for changing the spin axis direction in most real spacecraft, unless
the dissipation rate is very small.
Since the precession rate is constant, the precession angle will vary by
=

Js

t
Jt Js cos

After a time interval equal to


T =

(Jt Js ) cos
Js

the precession angle is varied by rad, and the direction of the symmetry axis of the
spinning satellite has changed of 2. If the firing time ton is such that the resulting
nutation angle is half of the desired variation of the spin axis direction and the thruster
activation time t0 is chosen in such a way that the resulting angular momentum increment
lies in the manoeuvre plane (that is, the plane containing both the initial and final position
of the satellite spin axis), then after T seconds the spin axis will be in the desired position,
At this point a second thruster firing is required
alhough moving at the coning speed .
to stop the precession and achieving again a pure spin condition about the new re-oriented
axis.
From simple trigonometric consideration, it is easy to show that
||h1 || = ||h2 || = h0 tan = Js tan
where h0 is the magnitude of the initial angular momentum. But being also
||h|| = Mt

42

G. Avanzini, Spaceflight Dynamics 2. Dynamics of Spinning Spacecraft

the required firing time for each pulse is equal to


ton =

Js tan
M

If is the thruster moment arm (i.e. the distance between the thruster axis and the
satellite centre of mass; in Fig. 2.9 this distance is equal to the radius of the spinning
body) and only one thruster is employed, it is M = F . In general at least two thrusters
are activated in order to obtain a balanced control torque, as stated before, in which case
it is M = 2F
In order to estimate the amount of propellant necessary to perform the reorientation,
it must be remembered that the specific impulse of an engine is given by
Isp =

F
mg/

that is the thrust delivered divided by the weight of propellant burned in a unit time.
The specific impulse is one of the most important characteristics of rocket engines of any
size, and from its value it is possible to determine
m =

F
gIsp

The total propellant mass required to rotate the spin axis of an angle equal to 2 is
mtot = 4

Js tan
gIsp

if two thrusters are employed for delivering the control torque M.


For small reorientation angles it is tan , and the propellant consumption is
roughly proportional to the reorientation angle itself. But as increases, the fuel consumption gets larger, so that it is more convenient to perform the overall reorientation
in several smaller steps. Moreover, for a given Isp , the pulse duration may become a not
negligible fraction of the spin period, thus reducing the change in angular momentum.
The drawback of a large reorientation performed in N steps is that the manoeuvre time
is increased and the overall manoeuvre is usually less accurate.
Figure 2.10 shows the results for reorientation 2 angles between 0 and 70 deg, for a
satellite with Js = 20 kg m2 , Jt = 15 kg m2 , spinning at = 2 rad s1 , that is controlled
by pair of thrusters with a 1 m moment arm, which produce 10 N of thrust each, with a
specific impulse Isp = 120 s.
The minimum pulse duration considered is 10 ms, that puts a limit on the resolution
of the reorientation, the minimum manoeuvre angle being 0.573 deg. On the other side,
a thruster firing time longer than one tenth of the spin period cannot be considered a
pulse. This limits the maximum reorientation angle that can be achieved by a single nutation manoeuvre, that is 17.9 deg. A stricter requirement on the pulse duration
will reduce the maximum available reorientation angle with a single manoeuvre. Using 2
steps, the propellant savings at low angles are negligible (less than 1%), but it is possible
to achieve spin axis reorientation up to 35 deg. With 4 steps the maximum reorientation angle becomes as high as 71 deg. Of course, the manoeuvre duration will increase
accordingly.
The correct timing of the thruster pulses is of very important in order to obtaine an
accurate reorientation. In this framework it should be noted that, unless a very particular

0.05
0.04
m

0.03
0.02
0.01
0
0

20

40

2 [deg]

man

=4

3
Nman = 2

2
1

Nman = 1
20

40

60

40

60

2 [deg]

400

single pulse t [ms]

Total firing time [s]

0
0

60

3
2.5
2
1.5
1
0.5
0
0

43

5
m tan

Manoeuvre time T [s]

Propellant mass m [kg]

G. Avanzini, Spaceflight Dynamics 2. Dynamics of Spinning Spacecraft

20

40

2 [deg]

60

300
200
100
0
0

20

2 [deg]

Figure 2.10: Spinning satellite manoeuvre: Fuel consumption (a), Total manoeuvre time
(b), Total firing time (c), single pulse duration (d).
case is dealt with, where the precession and the spin rate are in a rational ratio, the pair of
thrusters used for starting the manoeuvre will not be in the correct position for delivering
the required torque in the manoeuvre plane after half of the precession revolution. This
means that two pairs of thrusters need to be activated for the second pulse, delivering a
total net torque equal to the required one. As a consequence, the above description of
the manoeuvre technique will provide an estimate of fuel consumption slightly optimistic
and the overall complexity of the manoeuvre technique is, if possible, even higher in the
actual implementation.

44

G. Avanzini, Spaceflight Dynamics 2. Dynamics of Spinning Spacecraft

Chapter 3
Dualspin satellites
The presence of only one single inertially fixed direction on board of gyroscopically stabilized satellites greatly limits their use, as far as it is possible to mount a sensor payload or
a communication antenna only along that direction. Moreover, maneuvering the satellite
in order to reorient the spin axis is both difficult and expensive in terms of fuel consumption, the required propellant being proportional to the amount of angular momentum
stored in the spinning satellite.
In order to overcome this negative features, an alternative satellite configuration was
proposed and used in the past, where the payload and the communication hardware is
mounted on a platform P which has a relative rotational degree of freedom with respect
to the spinning part of the satellite, the rotor R. During orbit injection the two elements
are locked and the satellite behaves like a standard rigid body. After the operational
orbit is reached, the spinup motor accelerates the rotor, with respect to the platform,
in order to transfer on the rotor the entire vehicle angular momentum, thus despinning
the platform. The despin maneuver is ended when the platform is fixed in space, while
the rotor provides gyroscopic stability to the entire vehicle. The main advantage of such
a solution is that, pivoting the payload mounted on the platform allows aiming it in any
direction in space, with great flexibility, without changing the direction of the rotor spin
axis.
From the configuration point of view, the first dualspin satellites were made of a
large rotor, containing most of the satellite equipment, with solar cells mounted on the
surface of the rotor. The configuration resembled that of a standard (for that time)
spinning satellite. A small platform was attached to it, as represented in Fig. 3.1. Later,
the platform became larger (about one half of the satellite). The Intelsat IV was the
first prolate dual spin satellite (3.2). Finally, the rotor was substituted by a momentum
wheel, which is a small (and light) wheel spinning at a very high speed, placed inside the
platform. In this latter case, the satellite is called a momentumbias satellite.
The mathematical model that will be derived in the next section is not affected by
the relative size of platform and rotor. In this framework, both configurations (the dual
spinner, with a large rotor attached outside of the platform, and the momentum bias
satellite, with a small spinwheel inside the spacecraft bus) can be referred to as gyrostats.
The presence of a relative rotational degree of freedom between the platform and the rotor
(or spinwheel) of the gyrostat makes it necessary to derive a new formulation for the
angular momentum balance, since the body, featuring a rotating element, is no longer
rigid.
45

46

G. Avanzini, Spaceflight Dynamics 3. Dualspin satellites

Figure 3.1: Dual spin satellite: configuration with large rotor (e.g. GOES-7 satellite).

Figure 3.2: Dual spin satellite: configuration with equivalent rotor and platform (e.g.
Intelsat IV satellite).

G. Avanzini, Spaceflight Dynamics 3. Dualspin satellites

47

3
e

2
e
O

1
e
P

Figure 3.3: Sketch of a satellite equipped with a momentum wheel.

3.1

Mathematical model of a gyrostat

Figure 3.3 shows the generic arrangement of a momentum wheel W spinning about the
at an angular rate with respect to the satellite platform P. The mathematical
spin axis a
model is derived under the assumption that both the platform and the wheel are perfectly
rigid. Moreover, a perfectly balanced and axisymmetric wheel is assumed, the centre of
mass of which lies on the spin axis. As a result, the presence of a relative degree of
freedom between W and P does not affect the mass distribution of the spacecraft, which
. For this reason it is still
is independent of the angular displacement of W about a
possible to define a body frame FB , fixed with respect to the platform, the inertia tensor
I of the gyrostat P + W being constant in FB as for a standard rigid body. This means
that I represents the total moment of inertia of the gyrostat, including the wheel.

3.1.1

Angular momentum formulation

If the wheel is locked to the platform, the angular momentum is given (as usual) by
~ = I
~ , but if the wheel is spinning, than the overall angular momentum features a
h
second term,
~ = I
~ + Is
h
a
, the quantity hr = Is is the
where, letting Is be the moment of inertia of W about a
relative angular momentum. The absolute angular momentum stored in the spinning
rotor

~
ha = hr + Is a
is made up of two contributions, the relative angular momentum due to the spin motion
with respect to P, and the angular momentum related to the motion of the platform.

48

G. Avanzini, Spaceflight Dynamics 3. Dualspin satellites

Writing the components in body frame, it is


hB = I B + Is aB

ha = Is + aTB B

The equation of motion of the gyrostat can be written in terms of evolution of the angular
momentum vector as
h B = B hB + M B
h a = ga
where M B is the external torque acting on the satellite, while ga is the torque applied by
the despin engine to the rotor. It should be noted how ga , which is an internal torque,
does not appear in the equation of the time evolution of the overall angular momentum.
Nonetheless its effects on the overall dynamics are significant.
Also, the presence of the additional rotational degree of freedom increases the order
of the system of first order nonlinear differential equations with respect to the rigidbody
case, from 3 to 4 (three components for the total angular momentum plus the absolute
wheel angular momentum about the spin axis). In order to integrate this fourthorder system and to evaluate the Euler angle rates, it is necessary to determine the angular velocity
B . Integrating the Euler angle rate equation (or the quaternion evolution equation), one
gets the time evolution of satellite attitude.
The kinematic problem for the attitude of the satellite is solved expressing B as a
function of the total and wheel angular momentum, that is
B = I 1 (hB Is aB )
The value of B thus depends on both hB and . The rotor angular speed can be
expressed as
= ha /Is aTB B
where B appears, so that the two equations must be solved simultaneously, but this can
be done easily, since the two equations are linear in the angular velocity components. If
one assumes a set of principal axes of inertia, such that I = diag(Jx , Jy , Jz ), the resulting
linear system is given by

h1
1
Jx
0
0 Is a1

0
Jy
0 Is a2
2 = h2

0
h3
3
0
Jz Is a3
ha

Is a1 Is a2 Is a3 Is
Note that the 4 4 matrix with the moments of inertia must be inverted only once.

3.1.2

Angular velocity formulation

An alternative formulation of the equations of motion is also available, where gyrostat


dynamics is expressed directly in terms of the angular velocity components and rotor spin
rate. In this case, the angular momentum vector is writte as

hB = I B + ha Is aTB B aB

= I Is aB aTB B + ha aB

G. Avanzini, Spaceflight Dynamics 3. Dualspin satellites

49

The time derivative of the angular momentum vector components are thus expressed as

h B = I Is aB aTB B + ga aB

At the same time it is



h a = ga = Is + aTB B

so that, upon substitution of h B and h a into the angular momentum formulation, the
equations of motion of the gyrostat become
1
B = I [M B B (I B + Is aB ) ga aB ]
= ga /Is aTB B

where I = I Is aB aTB is the pseudoinertia tensor. These equations of motion can
be used for describing the gyrostat spinup dynamics, during platform despin, that is,
a spin axis torque ga is applied until the platform angular velocity drops to zero. The
knowledge of B allows for a direct evaluation of the Euler angle rates.

3.2

Simplified models

3.2.1

The Kelvin gyrostat

In many application, especially once the despin manoeuvre is completed, it is possible


to assume that the rotor relative spin rate remains constant, under the action of a control
action that drives the motor torque in order to compensate for fluctuations.
If one assumes that the angular speed of the rotor remains exactly constant, the
system looses one degree of freedom, as (t) = const and the associated dynamics can be
neglected. In this case the socalled Kelvin gyrostat is obtained. The variable hr = Is
becomes a system parameter and the platform equation of motion decouples from rotor
spin dynamics. In such a case, it is
I B = B (I B + hr aB ) + M B

In order to keep constant (that is, Omega


= 0), it is necessary to apply a certain spin
torque to the rotor, which, from the equation of motion of rotor spin dynamics can be
evaluated as
ga = Is aTB B
In real applications it is not possible to keep the rotor speed exactly constant and a
simple feedback loop is employed for determining the required spin torque. A tachometer
is mounted on the wheel shaft, which measure the actual rotor spin rate . The error
with respect to the nominal (desired) rotor speed des is the input for the rpm regulator.
As an example, the spin torque can be set equal to
ga = K(des )
where K is the gain of the rotor control system. In this way the rotor is accelerated
(ga > 0) whenever the angular speed is lower than the prescribed one, and decelerated
when > des . Of course, if the spin wheel dynamics is modeled including this simple
control scheme, it is no longer possible to treat as a constant in the platform angular
velocity equations, and the full system equations must be accounted for.

50

3.2.2

G. Avanzini, Spaceflight Dynamics 3. Dualspin satellites

The apparent gyrostat

If the rotor spinup torque is zero (ga = h a = 0), the rotor absolute angular momentum becomes constant and the apparent gyrostat model is dealt with. Also in this case
one state variable is lost, since it can be determined from the knowledge of a constant
parameter, namely ha . The equation of motion can thus be rewritten as


B + ha aB + M B
I B = B I

with the usual meaning for the symbols.


Note that, provided that the spacecraft moment of inertia, I, and wheel relative
and the absolute
angular momentum, hr , are replaced by the pseudoinertia matrix, I,
angular momentum, ha , respectively, the two reduced order models for the Kelvin and
the apparent gyrostats share the same expression.

3.3

Stability of axial gyrostat

Although of little practical use, the apparent gyrostat model is interesting for the derivation of analytical results on the stability of the despun condition. The simple axial case
is discussed in this paragraph.
Assuming that FB is a set of principal axes of inertia and that the spin axis is parallel
3 (axial rotor), the satellite and rotor absolute angular momentum can be written,
to e
respectively, as

Jx 1

hB =
Jy 2
; ha = Is ( + 3 )

Jz 3 + Is

The equation of motion of the axial gyrostat can be written as follows:


Jx 1 + (Jz Jy )2 3 + Is 2
Jy 2 + (Jx Jz )1 3 Is 1
Jz 3 + (Jy Jx )1 2 + Is
Is ( + 3 )

=
=
=
=

M1
M2
M3
ga

When = 0, the dualspin satellite behaves like a standard rigid body. This situation
is called all-spun condition. After platform despin, the satellite angular velocity must be
zero and all the satellite angular momentum must be stored in the rotor. The stability
of the equilibrium for the despun condition of a Kelvin gyrostat with axial rotor can be
analyzed as usual assuming the components B = (1 , 2 , 3 ) as small and neglecting
second order terms in the equation of motion, which can be written for the torquefree
case as
Jx 1 + Is 2 = 0
Jy 2 Is 1 = 0
Jz 3 = 0
3
The third equation is decoupled from the others, stating that a perturbation about e
will remain constant. The first and the second equations can be recast in matrix form as

 


1
0
(Is /Jx )
1
=
2
(Is /Jy )
0
2

G. Avanzini, Spaceflight Dynamics 3. Dualspin satellites

51

It is easy to show that the eigenvalues of the state matrix are a pair of conjugate
imaginary numbers, provided that 6= 0. No unstable eigenvalues are present for any
value of , so that exponential divergence is ruled out. It is possible to show, on angular
momentum conservation grounds, that the variation of 1 and 2 remains confined in a
neighborhood of the origin, so that the system is stable in the sense of Lyapunov. This
means that the rotor provides the required gyroscopic stability to the entire satellite.
Adding damping (e.g. under the action of a nutation damper) it is possible to transform the L-stable despun condition into an asymptotically stable equilibrium.

52

G. Avanzini, Spaceflight Dynamics 3. Dualspin satellites

Chapter 4
Active Stabilisation and Control of
Spacecraft
For many applications where accurate payload pointing is required full three-axis control
is used. The satellite does not spin (except during slew manoeuvres) and is stabilised by
reaction wheels (gyros) and/or thruster relative to a given desired attitude (e.g. an Earth
facing attitude). Since the satellite is not spinning, large Sun facing solar panels can
be used for efficient power generation. The satellite is also capable of fast reorientation
manoeuvres (slews) in any arbitrary direction, if sufficient control power is available from
the actuators. However, due to the added complexity, 3-axis stabilised satellites are
far more expensive than simple spinstabilised ones. The application and its accuracy
requirements must justify the increase in costs and the reduced life expectation of the
vehicle.
In order to actively stabilize the satellite and allow for autonomous maneuvering, the
satellite must be equipped with a suitable set of sensors, which provide the onboard
computer with the necessary information on the current attitude. The precision of the
control action depends on the accuracy of the attitude measurement. Moreover, the
control law itself depends on the type of actuators available. As a consequence, a brief
outline on the more common control actuator is given, prior to the discussion of attitude
active control.

4.1

Actuators

Spacecraft control actuators can be based on different physical principles. The choice in
the type of actuator installed on board of the satellite greatly affects its configuration, the
resulting pointing accuracy, and even the implementation of the control laws themselves.
A control action can be delivered to the satellite by either a direct torque produced
by a (set of) thruster(s), by exchanging angular momentum between parts in relative
rotational motion or by application of a magnetic torque, obtained from the interaction
of an electric coil and the Earth magnetic field.
Reaction control system (RCS)
Cold gas jets
Chemical propulsion (Monopropellant or Bipropellant propulsion system;
can be used also for orbit control)
53

54

G. Avanzini, Spaceflight Dynamics 4. Active Stabilisation and Control of Spacecraft

Momentum exchange devices


Reaction wheels
Biasmomentum wheel
Doublegimbal biasmomentum wheel
Control Moment Gyroscopes
Magnetotorquers

4.2

Linear model of rigid satellite attitude motion

For a general three axis motion, Eulers equation states that


h B + B hB = M B

(4.1)

If there are no internal moving parts (e.g. no spinning rotors) this equation can be written
in the form
I B + B (I B ) = M B
Assuming that the rotation is slow, the angular velocity can be considered as a first order
perturbation of the steady state B = 0, so that ||B || 1 and B (I B ) is a second
order (negligible) term in the equations of motion. Eulers equation thus becomes
I B = M B
When dealing with a stabilization problem, that is small perturbations with respect to
a prescribed nominal attitude are considered, also the angular displacement is small. In
such a case, it is possible to demonstrate that, letting the variation of the Bryans angles
be equal respectively to
= 3 ; = 2 ; = 1
and defining = {1 , 2 , 3 }T , it is

= B

Eulers equation in linear form become simply


= MB
I

(4.2)

Finally, assuming that a set of principal axes of inertia is chosen as the body frame, the
inertia matrix is in diagonal form, I = diag(J1 , J2 , J3 ), and three independent second
order linear ordinary differential equations are obtained, one for each axis,
Ji i = Mi , i = 1, 2, 3
It should be noted that for single axis rotations, a scalar equation in the same form as
Eq. (4.2) is obtained, regardless of the angular speed and angular displacement, that is
J = M
This equation will be the starting point for the analysis of attitude slew maneuvers.

G. Avanzini, Spaceflight Dynamics 4. Active Stabilisation and Control of Spacecraft

4.3

55

Linear model of gyrostat attitude motion

If the satellite is equipped with a momentum wheel with a moment of inertia Is spinning
with a relative spin rate , the total satellite angular momentum is
about the axis a
hB = I B + hr aB
where hr = Is is the relative angular momentum; can be written as
= 0 +
being 0 , i.e. is a small perturbation of the reference wheel spin rate 0 .
Equation (4.1) thus becomes
B + B [I B + Is (0 + ) aB ] = M B
I B + Is a
Neglecting higher order term one gets
B + Is 0 B aB = M B
I B + Is a
The evolution of the perturbations of Eulers angles is again
= B
so that the equations of motion becomes can be expressed in scalar form as follows:
1 + Is 0 (2 a3 3 a2 ) = M1
J1 1 + Is a
2 + Is 0 (3 a1 1 a3 ) = M2
J2 2 + Is a
3 + Is 0 (1 a2 2 a1 ) = M3
J3 3 + Is a
This set of equations is coupled with the dynamics of the spinwheel, described by the
(linear) equation
ga
TB aB
=
Is
i ), the
When the rotor spin axis is parallel to the ith principal axis of inertia (
ae
ith equation of motion becomes decoupled from the others, because aj = ak = 0, for
e
2 , so that a1 = a3 = 0 and a2 = 1. The
j, k 6= i. As an example, let us assume that a
equations of motion become
J1 1 Is 0 3 = M1
J2 2 + Is = M2
J3 3 + Is 0 1 = M3
The pitch dynamics is decoupled from the roll and yaw motions, but is coupled with the
spinwheel dynamics
ga
=
2
Is
Summing up, the torquefree pitch dynamics (M2 = 0) is described by the equations
(J2 Is ) = ga
= 0

J2

Is

On the converse, roll and yaw motions are coupled by the gyroscopic term. Moreover,
if the rotor spin speed is timevarying and is not negligible, the roll and yaw coupling
terms depend on the pitch dynamics through .

56

G. Avanzini, Spaceflight Dynamics 4. Active Stabilisation and Control of Spacecraft

4.4
4.4.1

Use of thrusters for attitude control


Single axis slews (open loop)

If we consider a constant torque acting around the generis ith axis, the angular acceleration is
i = i = Mi /Ji
0 ) = 0 (and dropping the subscript
Integrating from the initial condition (t0 ) = 0 , (t
i), one has

(t)
= 0 + (t t0 )

1
(t) = 0 + 0 (t t0 ) + (t t0 )2
2
By application of the
The same equation can be integrated in the phase plane vs .
derivation chain rule it is
d d
=
d dt
so that one gets
d
=
d
Integrating the previous equation in , one gets
Z

=
d

that, for constant applied torque, becomes

or,

1  2 2 
0 = ( 0 )
2
2 = 02 + 2( 0 )

Starting from a rest position in the nominal attitude ( = 0 and = 0), one gets

= 2
2 = 2 = (t)
A singleaxis rotation manoeuvre can thus be performed in three phases: an initial
acceleration phase (i) during which the spacecraft achieves the desired slew rate d for the
reorientation; the acceleration is followed by a drifting phase (ii), during which a constant
angular velocity remains constant with no action from the thrusters, until, at a proper
time instant, the drifting motion is stopped during the deceleration phase (iii) by a second
impulse from the thrusters. Representing the manoeuvre in the phase plane, the state
variables evolve as reported in Fig. 4.1.
The configuration of the thrusters used for spinning the satellite up to the drift angular
~
velocity can be different. As reported in Fig. 4.2, a single thruster delivering a force F
with a moment arm with resepct to the spacecraft centre of mass produces a torque the
magnitude of which is F . Unfortunately, also a net force F is produced which will affect
orbit parameters, thus coupling the orbit and the attitude control proboems.

G. Avanzini, Spaceflight Dynamics 4. Active Stabilisation and Control of Spacecraft

57

acceleration

drift

deceleration

Figure 4.1: Evolution of and for a resttorest manoeuvre.



X
~
X
 - F



X
~
X
 - F


~
F




X
X


Figure 4.2: Unbalanced and balanced thruster configurations.


On the converse, if a balanced configuration is employed, the simultaneous firing of
two thrusters in opposite directions will produce only a torqe M = 2F , with a zero net
force, so that the orbit motion of the spacecraft is unaffected by attitude control pulses.
An approximate evaluation of the total manoeuvre time and the propellant necessary
for performing it can be derived by neglecting the duration of the thruster firings for
accelerating and decelerating the angular motion of the vehicle, taking into account only
the drift phase. Assuming that the drift angular velocity is d , the time required for a
manoeuvre of amplitude equal to f is simply given by T = f /d .
Assuming a sharp thrust profile, it is
2F
d =
t
J
and the pulse width (thruster ontime) for the spinup is given by
ton,1 =

J d
2F

58

G. Avanzini, Spaceflight Dynamics 4. Active Stabilisation and Control of Spacecraft

From the definition of the thruster specific impulse


Isp =

F ton
gmf

(that is the ratio between the momentum gained and the weight of the propellant used)
the mass of propellant necessary for the acceleration is
m1 = 2

F ton,1
J d
=
gIsp
gIsp

The drift angular velocity can be determine as a function of the manoeuvre time, that
is, d = f /T , so that the propellant mass necessary for the spinup becomes equal to
m1 =

Jf
gIspT

At the end of the drift interval, a second pulse of equal duration in the opposite direction
will stop the motion of the satellite and the new attitude will be acquired. The total
propellant mass is thus equal to
Jf
mtot = m1 + m2 = 2
gIspT
It can be observed that allowing larger manoeuvre times, the same angle variation
can be achieved with a significantly smaller amount of fuel. Moreover, low energy thrust
pulses are less likely to excite vibrations in the structure and/or in flexible appendages
attached to the spacecraft (such as antennas or solar panels). For large values of T , smaller
ontime for the thruster are required, and the initial assumption that the acceleration and
deceleration phases have a negligible duration with respect to the overall manoeuvre time
is confirmed.
On the other side, for agile spacecraft that achieve high angular velocities, it is possible to improve the accuracy of the estimate of manoeuvre time taking into account the
duration of the initial and terminal phases. The time for accelerating the satellite up to
d is again given by
J d
d
=
t1 =
F

which is still equal to the time t2 necessary for bringing it back to rest at the end of the
maneuver, with the difference that for fast manoeuvres the total firing time t1 + t2 is
a not negligible fraction of the total manoeuvre time T .
During the acceleration, the variation of is
1
1 d2
1 = t21 =
2
2
and the total variation during the manoeuvre is given by
f = d td + 1 + 2
where the angle variation during the acceleration and deceleration phases, 1 and 2 , are
equal. The drift time is td = T 2t1 , where T is the total manoeuvre time, so that
f = d (T 2t1 ) + t21
!
d

2
= d T 2
+ d

G. Avanzini, Spaceflight Dynamics 4. Active Stabilisation and Control of Spacecraft

59

The total manoevure time becomes thus


T =

f
d
+

In the limit, the minimumtime bangbang control law brings to zero the drift time
td . The reader can demonstrate that in this limit case it is
q
T = 2 f /
In both cases, the faster manoeuvres are performed at the expenses of a significant
increase in propellant consumption.

4.4.2

Closedloop control

The ideal case: thrust modulation


If the satellite is equipped with a set of sensors that provides the necessary information
on the attitude motion, it is possible to implement a control law which provides a torque
command that drives the satellite towards the prescribed attitude.
If the torque command is simply proportional to the attitude error, e = des , the
closedloop equation of motion becomes

K
= M/J
= = (des ) = + p2 = p2 des
M = K(des ) = Ke
J
where p2 = K/J. The solution of this second order equation for an initial condition of
rest for = 0 is
(t) = des [1 cos (pt)]

i.e. an unacceptable undamped oscillation of amplitude des about the desired condition.
In order to damp the oscillation and asymptotically reach the desired condition, it is

necessary to add a damping term in the control law, proportional to the angular rate .
The command torque is thus
M = K(des ) + Kd

(4.3)

and the closedloop equation of motion becomes


+ c + p2 = p2 des
where the coefficient of the damping term c = 2p = Kd /J is positive (damped oscillations) if the gain associated to the angular rate is negative.
There are two possible resulting closedloop behaviours. For 0 < < 1, the time
history of the angular motion is
(t) = des [1 exp(pt) cos (pd t)]

p
where pd = p 1 2, that is, damped oscillations with an exponentially decaying amplitude.
If onpthe converse > 1, the characteristic equation has two real solutions 1,2 =
p( 2 1), which are both negative. In this case the evolution of the rotation
angle is


2
1
(t) = des 1
exp(1 t) +
exp(2 t)
2 1
2 1

60

G. Avanzini, Spaceflight Dynamics 4. Active Stabilisation and Control of Spacecraft

Figure 4.3: PulseWidth/PulseFrequency (PWPF) modulator.


Two important issues need to be considered, at this point. The first one is the choice
of the gains. The case with < 0 is not considered at all, inasmuch as one of the poles of
the closed loop system would be a positive real number, that is, the closed loop system
would be unstable. When < 1 (subcritical damping), the time to damp out 99% of
the initial oscillation amplitude increases as gets smaller according to the equation
exp(p ) = 0.01 = = log(0.01)/(p)
At the same time, if > 1 (supercritical damping) the time constant of the slowest mode
becomes larger, being approximately
equal to /p, for large values of . This means that
2
1
1
a largely
supercritical damping
coefficient
will cause a slower convergence.
.5s+1
Input
Output
The best performance in terms of maneuver agility are obtained for the critical dampModulator
ing = 1, which guarantees the
filter fastest convergence time to the desired position. The
Schmidt
trigger
overdamped case may be of interest in those cases
when
fuel consumption is more a
stringent concern than maneuver time. On the contrary, the underdamped case, which
exhibits some overshoot, is of no practical interest, since it is characterized by a longer
settling time and an increased fuel consumption, because of the thrusting system firing
alternatively in both directions, during the oscillations.
The second issue concerns the actual implementation of a control law in the form (4.3).
Such a control law would be realistic only if the control torque provided by the thruster
could be modulated. Unfortunately direct thrust modulation would require a complex
hardware and, at the same time, would be extremely inefficient in large portions of the
required thrust range. For these reason, thruster are always on/off devices. As a consequence, some form of pulse modulation is required to provide a feasible implementation of
the control torque demand. Two techniques will be presented, the PulseWidth/Pulse
Frequency (PWPF) modulator, which is an analogic device, and the PulseWidth Modulator (PWM), which lends itself to a discrete time implementation.
PulseWidth/PulseFrequency modulation
Figure 4.3 shows the structure of a PulseWidth/PulseFrequency (PWPF) modulator.
The main element of the modulator is the Schmidt trigger, which consist of a double relay
with hysteresis, separated by a dead band. In order to provide a quasilinear steadystate
response, a modulation filter is added, the input of which is the signal e = em ym , i.e.
the difference between the feedback signal error and the modulator output. The output
of the filter v is the activation signal for the Schmidt trigger.

G. Avanzini, Spaceflight Dynamics 4. Active Stabilisation and Control of Spacecraft

61

The modulator filter is represented by a linear first order system, the dynamics of
which is
1
v = (Kem v)

The response of the modulator filter to a steady input e is


v(t) = v0 exp (t/ ) + Kem [1 exp (t/ )]
until y = 0.
The output y of the modulator remains zero until the signal v remains below the
activation threshold Uon . This threshold is never crossed if the error variable em remains
smaller than Uon /K, that is, the asymptote of v(t) Kem lies below Uon .
If on the converse Kem > Uon , there will be a time tom when v reaches Uon and the
trigger output switches from 0 to Um (Fig. 4.4), so that the thruster is activated by the
trigger. At this point, the modulator input changes to e = em Um . Assuming that em is
unaffected by y, integration of the filter linear dynamics starting from the initial condition
v(ton ) = Uon for e = em Um brings to the following expression for the activation function





t ton
t ton
+ K(em Um ) 1 exp
v(t) = Uon exp



t ton
= K(em Um ) + [Uon K(em Um )] exp

The thruster is switched off when the activation function becomes equal to Uoff , that is
when


toff ton
Uoff K(em Um )
exp
=

Uon K(em Um )
Uon Uoff
= 1
(4.4)
Uon K(em Um )
It should be noted that, if K(em Um ) > Uoff , i.e. em > Uoff /K + Um , the above equation
has no real solution, that is the input is so great that the modulator calls for a continuous
thruster engagement. This level constitutes the saturation level of the modulator. On
the converse, if em becomes less than Uoff /K + Um , the thrusters are switched off when
t = toff . Calling ton = toff ton the ontime of the thruster, and assuming that it is
sufficiently small, it is possible to approximate exp (ton / ) = 1 ton / , so that
ton

Uon Uoff
Uon K(em Um )

After the thruster pulse, the modulator input becomes again e = em and the response
of the filter for an initial condition given by v(toff ) = Uoff is





t toff
t toff
v(t) = Uoff exp
+ Kem 1 exp

The thruster will be activated again when v = Uon , that is when




Uon Kem
t toff
exp
=

Uoff Kem

62

G. Avanzini, Spaceflight Dynamics 4. Active Stabilisation and Control of Spacecraft

1.2
first activation when V = Uon

Filter variable

Uon

0.8
0.6
0.4
U

0.2
0

off

0.2

0.4

0.6

0.8

1.2

1.4

1.6

1.8

1.2

1.4

1.6

1.8

deactivation when V = U

off

initial condition V = 0

Output torque

10
pulse
duration
t

on

6
4
time between
pulses t

off

0.2

0.4

0.6

0.8

1
time [sec]

Figure 4.4: Activation/deactivation sequence of a PWPF modulator.


Assuming again that the off time is also sufficiently small, one gets
toff =

Uon Uoff
Kem Uoff

Figure 4.5 shows the activation sequence of a PWPF modulator for increasingly higher
values of the error signal em . In the first case it is clear how the error signal is so small
that the thruster are not activated. For an error sligthly above the activation threshold,
a sequence of small pulses (25 ms) separated by wide deactivation intervals (toff 0.6
s) are present. This correspond to a small average control torque. A higher error signal
increase the duration of the pulses, while decreasing the time between the pulses, so that
the average torque increases almost linearly until the saturation level is reached. Beyond
saturation (bottomright corner in Figure 4.5) the deactivation threshold is no longer
crossed and the thruster are not switched off.
The average output of the thruster, given by
y = Um

ton
toff + ton

is represented in Fig. 4.6. Its variation is almost linear for Uon /K < em < Uoff /K + Um , it
is zero below the activation threshold and equal to the thruster force beyond the saturation
value. The proper sizing of the trigger and filter characteristics through tuning of filter and
trigger parameters allows for realizing a modulator which takes into account the physical
characteristics of the thrusters, such as the minimum pulse time and the minimum time
between pulses.

G. Avanzini, Spaceflight Dynamics 4. Active Stabilisation and Control of Spacecraft

1
Filter variable

Filter variable

1
0.8
0.6
0.4
0.2
0

0.5

1.5

Output torque

Output torque

4
2
0

0.5

1
time [sec]

1.5

0.5

1.5

0.5

1
time [sec]

1.5

0.5

1.5

0.5

1
time [sec]

1.5

6
4
2

1
Filter variable

1
Filter variable

0.8
0.6
0.4
0.2

0.8
0.6
0.4
0.2

0.5

1.5

10
Output torque

10
Output torque

0.4

10

8
6
4
2
0

0.6

0.8

0.2
0

10

63

0.5

1
time [sec]

1.5

8
6
4
2
0

Figure 4.5: Activation sequences of a PWPF modulator for different input levels.

DC
1.0

0.5

0.0
0

e
Figure 4.6: Duty cycle of the PWPF modulator.

10

G. Avanzini, Spaceflight Dynamics 4. Active Stabilisation and Control of Spacecraft

Angle [deg]

Ang. speed [deg/s]

Torque [N m]

64

1
0
1
0

20

40

60

80

100

20

40

60

80

100

20

40

60

80

100

1
0.5
0

20
10
0

time [sec]

Figure 4.7: Closedloop behaviour (short term).


Figure 4.7 shows the response of a satellite controlled by a set of thruster driven by a
PWPF modulator (continuous blue line). The response is compared with that obtained
by thrust modulation (red dashed line). In the initial part of the reorientation manoeuvre
(t < 100 s), the modulator reproduce almost exactly the behaviour of the ideal linear PD
controller.
In the following part of the simulation, shown in Fig. 4.8, the angular velocity of
the ideal case goes asymptotically towards 0, while the small residual angular velocity
(-0.002 deg/s) visible in the more realistic case, where the modulator is present, causes a
slow drift towards the negative side of the deadband. Since the error signal is within the
deadband no control action is produced, until for t = 273.36 s a pulse is fired which revert
the motion. The minimum pulse provides the required reversal, but the resulting angular
velocity is significantly higher (0.025 deg/s) so that after few seconds, t = 286.43 s, a new
pulse in the opposite direction is required to revert again the motion on the positive side
of the deadband. This bouncing between the edges of the deadband characterizes the
steadystate condition of a spacecraft controlled with a reaction control system.
PulseWidth Modulation
The PWPF modulator is a heritage of the analogic era, although it is still widely used. The
main advantage of PWPF modulation remains the possibility of commanding a sequence
of pulses with a quasilinear response, in terms of averaged output. Nonetheless, many
actuators are now commanded in the framework of a digital implementation of the control
logic. It is well known that a digital controller obtained from the discretization of an
analogic one can attain at most the performance of the original analogic counterpart. If a
digital controller is designed directly in the discrete time domain, taking into account the
sampling frequency and A/D and D/A conversions, better performance can be obtained.

Ang. speed [deg/s]

Torque [N m]

G. Avanzini, Spaceflight Dynamics 4. Active Stabilisation and Control of Spacecraft

65

1
0
1
100

150

200

250

300

350

400

450

500

150

200

250

300

350

400

450

500

150

200

250

300
time [sec]

350

400

450

500

0.1
0
0.1
100

Angle [deg]

21
20
19
100

Figure 4.8: Closedloop behaviour (long term).


PulseWidth modulation provide a means for designing a thruster control law where
the implementation of the switching logic is inherently digital. As an advantage with
respect to PWPF (analogic) modulation, it is much easier to satisfy restrictions on the
ontime ton and the timebetweenpulses toff .
Figure 4.9 shows a sequence of pulses. If ts is the sampling period of the digital
controller, a pulse can be commanded for each duty cycle, the duration of which is tdc =
Nts . The minimum impulse bit (MIB) Non,min is chosen so that Non,min ts > ton,min .
At saturation, the ontime is extended to the entire duty cycle. Below saturation, the
maximum pulse must satisfy the constraint on minimumtimebetweenpulses (MTBP),
that is toff = (N Non )ts >MTBP.
The average output of a thruster controlled in PWM is
y = Um

ton
Non
= Um
toff + ton
N

The response of the PWM to a constant input error signal e is depicted in Fig. 4.10,
assuming that a quasilinear average output y = k
e is desired between a deadband edb
and the saturation value esat , when the ontime equals the duty cycle and y = Um . It
should be noted that, by a proper software implementation of the activation signal, it is
possible to obtain arbitrary pulsewidth modulation curves.

4.4.3

Fine pointing control

After slew to target, the control system switches to fine pointing control, in order to
maintain the payload aimed at the desired target with a prescribed tolerance, in presence
of external disturbances. This can be done using momentum exchange devices and/or gas
jets. A simple way to achieve such a fine pointing is to use cold gas jets, that produce

66

G. Avanzini, Spaceflight Dynamics 4. Active Stabilisation and Control of Spacecraft

y 6
Um


-

 -

ts

ton


toff
tdc

Figure 4.9: PulseWidth Modulation.


y
Um
1

0.8

0.6

0.4

0.2

0
0

0.2

0.4

0.6

0.8

1.2 e

esat

Modulator data: ts = 10 ms; Ndc = 20; ton,min = 12 ms Non,min = 2;


MTBP = 17 ms Non,max = 18.
Figure 4.10: Response of a PW Modulator.

thrust in the range between 0.05 and 20 N, with short pulse times (of the order of 102 s)
for fine control. As far as the satellite will be bouncing back and forth between the edges
of a limit cycle, it is important to estimate the amount of fuel necessary as a function of
the prescribed pointing accuracy and thruster characteristics.
Torquefree control
Let us consider the torquefree case, that is no external disturbance acts on the spacecraft,
where the payload must be pointed in a given direction, within some deadband, which in
the case of astronomy payloads can become as small as 1/10 arcsec = 0.000278 deg!
The satellite is allowed to drift slowly inside the deadband, with drift velocity d .
When it reaches the limit of the deadband, db , the thruster is fired so as to revert the
drift rate in the opposite direction. The satellite will drift towards the other side of the
deadband, where the thruster are fired again, thus capturing the satellite into a limit cycle
of width approximatley equal to 2db . In what follows it will be assumed that the limit
cycle is symmetric, that is, the absolute value of the angular velocity during the drifting

G. Avanzini, Spaceflight Dynamics 4. Active Stabilisation and Control of Spacecraft

67

0.01

thruster firing

thruster firing

[deg s1]

drift

drift
deadband
0.01
3

[10

deg]

Figure 4.11: Limit cycle for gasjet fine pointing control.


phases is the same in both directions.1 The thrusters generates a moment
M = 2F = J
For a thruster pulse width ton , the change of the attitude rate is
F
ton = 2d
J
for a symmetric limit cycle, in order to exactly revert the satellite motion after the firing.
Remembering that
F ton
Isp =
gmf
one gets, for each thruster firing, a fuel consumption of
= 2

mf = 2

F ton
J d
=2
gIsp
gIsp

Considering both sides of the deadband, the total fuel consumption per cycle is
mf,tot = 4

J d
gIsp

Ignoring the (usually negligible) thruster pulse time, the duration of the limit cycle is
TLC 4
1

db
d

As it was evident from the simulation in the previous paragraph, this is not what happens in reality,
since there is no way of enforcing a perfectly symmetric limit cycle with a simple PD controller, unless
more complex control strategies are implemented. Nonetheless, the following analysis is conservaive,
inasmuch as the symmetric limit cycle is the fastes one (its period is the shortest), thus requiring the
maximum number of pulses per unit time. This means that the estimate of the fuel consumption in this
situation is a worstcasescenario.

68

G. Avanzini, Spaceflight Dynamics 4. Active Stabilisation and Control of Spacecraft

and the average fuel consumption (that is the fuel burned per unit time) is
2
f = mf,tot = J d
m
TLC
gIspdb
which can be rewritten as

2
f = (F ton )
m
gIspJdb

This latter equation shows clearly that in order to achieve a great accuracy (small value
of db ) without paying too a great penalty in terms of fuel consumption, the characteristics
f remains small onfly if high
of the thrusters are of paramount importance, inasmuch as m
specific impulse thrusters are employed, capable of delivering the force F for a very short
pulse time ton .
Bias torque compensation
In most cases, the payload must be aimed at a target (be it on Earth or in the open
space) in presence of external disturbances. If the resulting torque is roughly constant,
such as that due to the gravity gradient, it is possible to use this torque for reversing the
motion on one side of the deadband. In this case, the environmental torque will bring the
spacecraft drifting towards the top of the tolerance interval. At this point the thruster are
fired, to reverse the motion, that will be slowed down by the adverse disturbance torque.
Rather than allowing the spacecraft to reach the other side of the deadband and fire a
second thruster pulse, one allows the bias torque to slow and reverse the drift inside the
deadband interval. The limit cycle will be made of two parabolic arcs, one relative to the
thruster firing phase, similar to that of the previous case, that reverse the satellite motion
on one side of the deadband, and the second arc, relative to the slow deceleration and
motion reversal due to the external torque.
If the disturbance torque generates an angular acceleration
= D = MD /J
it is

(t)
= 0 + D t
(t) = 0 + 0 t + D t2 /2
where the subscript 0 indicates the initial condition at time t0 . From the first equation,
0 )/D , that, substituted in the second one, after some simple algebra
it is t = ((t)
gives
2 = 02 + 2D ((t) 0 )
If the initial condition is on the deadband limit, it is 0 = db . In order to keep the limit
cycle inside the dead band, the turning point at which drops to zero and reverts its sign
must be crossed on the other side of the deadband, when 0 = db . In this case, it is
2 = 0 = 02 + 2D (2db )
that is

J 02
db =
4MD

G. Avanzini, Spaceflight Dynamics 4. Active Stabilisation and Control of Spacecraft

69

0.01

[deg s1]

drift

thruster firing

deadband
0.01
3

[104 deg]

Figure 4.12: Limit cycle for gasjet fine pointing control with constant disturbance torque.
In order to close the limit cycle, the thruster pulse must be such that the total is
twice the drift angular velocity 0 at deadband crossing,
= 20 = 2F ton /J
Substituting the latter expression in the equation of db one gets
db =

(F ton )2
4MD J

and again the requirement on the thruster is to deliver a small duration pulse for keeping
the deadband small enough to satisfy fine pointing requirements.
Exercise
After determining the duration of the limit cycle in presence of a bias torque, where as
usual the firing time can be considered negligible, prove that the average fuel consumption
f = mf,tot /TLC is independent of the db , if the motion reversal under the action of
m
the disturbance torque exploits the whole deadband amplitude.

4.5

Momentum exchange devices for attitude control

Gas jets can deliver very high torque to manoeuvre a spacecraft and change its orientation,
but they have a limited life, due to the finite propellant mass budget available. For this
reason, it is important to rely on some different kind of actuators for stabilising and
controlling satellite attitude during normal operations, without using propellant.
Reaction wheels (RWs) and momentumbias wheels (MBWs) can exchange angular
momentum with the spacecraft bus using only electrical power, that is obtained from the

70

G. Avanzini, Spaceflight Dynamics 4. Active Stabilisation and Control of Spacecraft

solar panels for the whole lifetime of the satellite. Each wheel is attached to the satellite
structure trough an electric motor, that can be used to accelerate or decelerate the wheel,
relatively to the satellite. If we assume that initially both the satellite and the wheel
are still in the inertial frame (zero angular momentum initial condition), spinning up the
wheel in one direction will cause the bus to rotate in the opposite direction, because of
conservation of overall angular momentum.
i . The
Let us consider a single axis rotation about the principal axis of inertia e
i , including the wheel. The
spacecraft has a total principal moment of inertia Ji about e
= e
i , having a spin moment of inertia Is about a
and
wheel spins around the axis a
angular velocity , relative to the spacecraft bus. Its rotation does not change the actual
mass distribution, so that Ji is independent of the motion of the wheel.
i is
The overall angular momentum about e
hi = Ji i + Is
where the first term is the angular momentum of the satellite induced by the overall
rotation rate i , while the second one is the relative angular momentum stored in the
wheel.
Note that the absolute angular velocity of the wheel is + i with respect to the
inertial frame FI . If initially i = = 0, so that hi = 0, and taking into account that
the total angular momentum is conserved, inasmuch as the wheel is spun up or slowed
through the exchange of internal torques applied by the electric motor, one gets
i =

Is

Ji

(4.5)

Usually it is Is Ji , so that i , that is the angular velocity achieved by the


spacecraft is only a small fraction of the relative spin angular velocity of the wheel.
In order to manoeuvre the spacecraft about 3 axes, it is necessary to equip the satellite
with at least 3 RWs, one for each body axis. The cluster of RWs usually contains a
fourth spare wheel, the axis of which is oriented in such a way so as to exchange angular
momentum components to or from any of the three axes, if one of the main wheels fails.
One of the wheels can be used to despin the satellite after orbit acquisition. In such
a case, this wheel shall be oriented as the spin axis of the satellite and apogee kick rocket
motor stack and it is called a momentum wheel. Its angular momentum is unlikely to
drop back to zero and the control torques about its axis will be obtained accelerating and
decelerating the wheel.
The MBW will also provide a certain degree of gyroscopic stability, which is an interesting feature. At the same time, the presence of the momentum bias will couple the
motion about the other two axis because of the gyroscopic effect observed when deriving
the equations of motion of gyrostats. As a consequence, it is no loger possible to consider
separately the control problem about the principal axis of inertia. In spite of this fact,
only singleaxis rotations will be considered in what follows, for the sake of simplicity.

4.5.1

Openloop control with RWs

Assuming for simplicity torquefree motion, we will size a reaction wheel with respect to
requirements in single axis slew manoeuvre. Dropping the i subscript, the total angular
momentum about one of the principal axis is
h = hs + hw

G. Avanzini, Spaceflight Dynamics 4. Active Stabilisation and Control of Spacecraft


6T

71

h s

= f
t4

t3
t1
=0

t2

h s
Figure 4.13: Torque profile for a single axis slew manoeuvre.
where hs = (J Is ) is the spacecraft angular momentum (without the wheel), while
hw = Jsw ( + ) is the spin wheel absolute angular momentum. Since there are no
external toruqes, h = const and
h = h s + h w = 0 = h s = h w
where ga = h w is the motor torque applied to the spin wheel.
and the equation of
Note that, assuming single axis rotations, it is = = ,
motion can thus be written in the form
(J Is ) = ga
which has exactly the same form derived for the RCS case, with the only difference of the
sign in the control torque, which, in the case of RWs or MBWs is a reaction torque.
As a consequence, a similar openloop control technique can be used, where one pulse
is used for spinning up the reaction wheel and consequently the vehicle in the opposite
direction, achieving a certain drift angular velocity, according to Eq. (4.5), while a second
pulse will spin down both the spacecraft and the wheel, achieving at the end of the
manoeuvre the desired attitude f .
After the first pulse, the spacecraft gains an angular momentum equal to
hs (t2 t1 ) = (J Is ) = h w (t2 t1 ) = ga (t2 t1 )
But it is also

h s = (J Is ) = (J Is )

where = = ga /(J Is ) is the angular acceleration of the spacecraft. The evolution


of the angular displacement is thus given, as in the RCS case, by
1
(t2 t1 )2
2
1 ga
=
(t2 t1 )2
2 J Is

(t2 ) =

Note that simpler expressions can usually be employed for preliminary sizing purposes, by
remembering that usually it is J Is . On the other side, the value of the available torque
is usually much smaller than that obtained from thrusters, so that the pulse duration is
not negligible with respect to manoeuvre time.

72

G. Avanzini, Spaceflight Dynamics 4. Active Stabilisation and Control of Spacecraft

As an advantage of reaction and momentum bias wheels, the fact that electrical power
(and no fuel) is employed for accelerating and decelerating the wheel make it possible
to perform fast manoeuvers, without paying any penalty in terms of operational lifetime
with respect to slower ones.2

4.5.2

Sizing a reaction wheel for single axis slews

For a minimumtime rotation, a bangbang control must be used, so that t2 t3 and the
final rotation angle is
h s
f = 2(t2 ) =
(t2 t1 )2
J Is
But assuming t1 = 0, it is also tf = 2(t2 t1 ), for bangbang control, and one gets
f =

1 h s 2
t
4 J Is f

Because of conservation of angular momentum, we recall that h s = h w , so that the


maximum torque that the electric rotor must produce is
(J Is )f
ga,max = |h w | = 4
t2f
But for t = t2 , it is also
2 ) = hw (t2 ) = h w (t2 t1 ) = h w tf /2
hs (t2 ) = (J Is )(t
when the wheel achieves the maximum angular momentum. This means that the momentum capacity of the wheel must be
hmax
2(J Is )
w

2
tf

Numerical example
We require a spacecraft with J = 100 kg m2 to perform a 0.2 rad slew in 10 sec. This means
that the electric motor must apply torques in both direction equal to (neglecting the unknown
value of Is )
Jf 4 100 0.2
h w 4 2
= 0.8 Nm
102
tf
The maximum momentum stored in the RW is reached when half of the reorientation is completed, so that the RW capacity must be at least
hmax
2J
w

2
2 100 0.2
=
= 4 kg m2 s1
tf
10

Assuming a wheel inertia Jsw = 0.03 kg m2 , the maximum wheel angular rate with respect to
the satellite bus is
J
J hs
= =
Is
Is J Is
hw

= 133 rad s1 = 21.2rpm


Jsw
2

On the converse, a faster drifting phase requires more acceleration and deceleration from the reaction
control system, which means more fuel is burned for achieving the same final attitude. This means that
the satellite will run out of fuel sooner, if fast manoeuvres are performed insterad of slow ones.

G. Avanzini, Spaceflight Dynamics 4. Active Stabilisation and Control of Spacecraft

4.5.3

73

Closedloop control with RWs for single axis slews

With the usual meaning of the symbols, the equation of motion for singleaxis rotation is

d 
J + Is = Mext
dt
so that it is

J + Is = Mext

(4.6)

The reaction wheel dynamics is described by the first order equation





Is + = ga Is = ga Is

Substituting the latter expression in Eq. (4.6) and assuming for simplicity that the
external torque and the momentum bias are both zero, it is possible to rearrange the
dynamics equation as
(J Is ) = ga
By choosing a simple PD control in the form
ga = KP (des ) + KD

(4.7)

the closedloop dynamics is described by the second order differential equation


J + KD KP = KP des
where J = J Is is the moment of inertia of the platform without the spinwheel.3
By defining the closedloop system bandwidth as p2 = KP /J (with Kp strictly
negative for static stability) and a damping parameter such that 2p = KD /J (with
KD strictly positive for asymptotic stability), the equation of motion can be rewritten in
the form
+ 2p + p2 = p2 des
which can be solved analytically and clearly presents a nice way of choosing the gains KP
and KD .

4.5.4

Bias torque and reaction wheel saturation

Let us assume for simplicity that the satellite is subject to a constant disturbance torque
MD . It is easy to see that the angular displacement evolves according to the differential
equation
+ 2p + p2 = p2 des + MD /J
and asymptotically converge to the steady state
ss = des + MD /KP 6= des
This means that the use of the control law of Eq. (4.7) may not guarantee a sufficient
pointing accuracy. The steady state error can be made smaller by increasing the proportional gain KP , but it will never vanish. Moreover, increasing the gain will likely degrade
3

Note that the feedback control law duplicates exactly the control law employed for reaction control
systems, Eq. (4.3), but because ga is a reaction torque, the gains appears with their sign changed in the
equation of the closed loop system dynamics. Their values must be chosen accordingly.

74

G. Avanzini, Spaceflight Dynamics 4. Active Stabilisation and Control of Spacecraft

the performance of the real system, since the sensor noise will also be multiplied by a high
gain.
It is possible to asymptotically cancel the steady state error by adding an integral
term to the control law, i.e.
ga = KP (des ) + KD + KI y
where
y = des y =

(4.8)

(des )dt

If an integral term is introduced, a steadystate can be achieved only if y = 0, i.e. = des ,


with a value of y such that
KI y = MD ga,ss = MD
This means that if the spacecraft is subject to a constantly not null external torque, the
average value of which is not zero (aerodynamic torque, gravity gradient), the platform
can be maintained aimed at a fixed position only if an equivalent motor torque is delivered
to the reaction wheel. As a consequence, the wheel will absorb these disturbances by spin
up and fix the orientation, but its angular velocity will constantly increases. Should the
wheel reach the maximum spinning speed, the attitude control system is saturated.
In order to avoid such a problem, before reaching saturation it is necessary to dump
the angular momentum excess. This can be done simply braking the wheel and firing the
thrusters in the opposite direction. In this way the momentum accumulated in the wheel
is brought back to zero, because of the external torque produced by the thrusters. During
wheel desaturation the pointing accuracy of the control system is usually reduced.
The use of thrusters for despinning the wheels shows that also when using RWs for
attitude control it is necessary to carefully design the propellant mass budget. Thruster
firings will be also in this case the main source of fuel consumption during the station
keeping phase, and will be the main limit to the satellite operative life.

Chapter 5
Environmental torques and other
disturbances
Although small in terms of absolute value, environmental torques acting on a spacecraft
for a very long period of time can produce a sizable displacemtne of the satellite from the
desired attitude. A 3axis stabilised spacecraft featuring only a reaction control system is
continuously bouncing backandforth between the edges of the allowed deadband, under
the alternating action of a (small) disturbance torque and the quick reversal provided by
a (large) control pulse at one of the edges of the deadband.
If the spacecraft is either gyroscopically stabilised or is characterised by a considerable
momentum bias the process may take longer, since a sizable variation of a large value of the
angular momentum requires more time. In any case, sooner or later, some compensation
is required. Disturbances of different nature may affect spacecraft attitude, especially
whem flying a Low Earth Orbit. At the same time, internal sources of disturbances are
also present, which either affect the dynamic behaviour of the satellite with respect to
the ideal one, or directly affect sensor pointing precision because of oscillations induced
by flexible appendages.
Among many others, gravity gradient torque has the interesting property of stabilising
an Earth pointing attitude, provided that a certain allignment condition between orbit
and principal frame is satisfied. When the requirement on pointing accuracy is weak (of
the order of some tenths of a degree) this passive stabilisation technique, that does not
require fuel, may be an option.

5.1

Environmental torques

The major sources of external disturbances that may influence the dynamic behaviour of
a satellite are:
gravity gradient: the variation of gravity pull acting on different mass elements
of the spacecraft provide a small yet significant torque; in Low Earth Orbit (LEO)
this torque can stabilise an Earth pointing attitude, although the resulting motion
is not naturally damped;
atmospheric drag: in Low Earth Orbit (LEO) the air density is small but not
null; the centre of pressure is usually not aligned with the center of mass and this
fact will produce a torque acting on the spacecraft;
75

76

G. Avanzini, Spaceflight Dynamics 5. Environmental torques and other disturbances

magnetic field: the satellite, with its metallic parts and currents flowing in the
electric system produces a magnetic field that interacts with the magnetic field of
the Earth, exspecially in LEO; again a torque on the satellite is produced by this
interaction; this coupling between a magnetic field produced by the satellite and
the environmental magnetic field can be used beneficially to provide a simple and
effective means of attitude control; a magnetotorquer is made of a magnetic coil
(or a permanent magnet) the orientation of which inside the satellite bus can be
changed so as to produce a control torque; as an alternative, to coils at an angle can
be activated with different currents so as to provide an arbitrary torque in a plane;
no torque can be delivered in the direction of the magnetic field, so that magnetic
torques alone are not sufficient for threeaxis control;
solar radiation: the solar radiation produce a light pressure force on the satellite
parts exposed to the flow of nuclear particles emitted by the sun; again the force
offset will determine a small torque, that is insignificant in LEO, but is the main
disturbance at higher distances from the Earth, such as in GEostationary Orbit
(GEO).
A simple mathematical formulation is available only for the gravity gradient torque,
that will be discussed later in this chapter. Atmospheric and solar radiation torques
depend strongly on the shape of the spacecraft, and in particular on how the solar panels
and the spacecraft bus face the flow of the rarified gas moleculae or the stream of high
energy particles from the Sun. The magnetic torque, on the converse, depend on the
electric system architecture. As a consequence no general model is available for these
disturbances.

5.2

Internal disturbances

There are also some sources of internal disturbances, that can be due to the uncertainty
of some parameters or to some simplifying assumptions the effect of which is not null on
the actual spacecraft:
uncertain center of gravity: there can be an uncertainty on the position of the
center of gravity up to some centimeters (1 to 3 cm);
thruster misalgnment: the direction of the thruster can be different from the
design one up to 1 deg;
runrun thrust variation: there can be sizable variation (up to 5%) in the thrust
produced by a thruster in different firings;
fuel slosh: fuel sloshing in the propellant tanks causes both shift in the position
of the centre of mass and in the mass distribution (that is, affects the value of the
moments of inertia); moreover, as propellant is burned, a long term variation in
mean mass properties is also experienced by the satellite;
rotating parts: rotating parts inside the satellite (gyroscopes, reaction wheels, but
also tape recorders) produce microvibrations which in turn cause loss in pointing
accuracy;

G. Avanzini, Spaceflight Dynamics 5. Environmental torques and other disturbances

77

elastic modes of flexible appendages: several flexible items, such as antennas,


booms and solar arrays can be attached to the spacecraft bus, and their flexibility
can interact with the rigid body motion of the main platform, resonating at the
bending frequencies, if properly excited, with serious consequences for the pointing
accuracy.

5.3

Gravitygradient stabilization

Spin stabilization allows one to provide the satellite with gyroscopic stability, which can
be made asymptotic, by a proper use of dissipative devices. At the same time, some
significant disadvantages affects the operational use of spin stabilized satellites, that are
the presence of only one inertially fixed direction along which to point the payload;
the inefficient production of electric power, inasmuch as the entire surface of the
satellite must be covered with photovoltaic cells, only half of which faces the sun
during the spinning motion (with less than one fourth at an angle greater than 45
deg with respect to the direction of the sun rays);
the complexity, limited accuracy and cost in terms of fuel of attitude maneuvers for
reorientation of the spin axis.
A very interesting possibility for stabilizing an Earth facing satellite in low Earth
orbit is to exploit the gravitational torque that acts on any rigid body of finite size.
The differential gravitational acceleration across the satellite provides a small, but non
negligible torque. Simply put, the side of the satellite closest to the Earth experiences a
slightly greater gravitational acceleration that the opposite side. The resulting torque will
tend to align the satellite with the local vertical direction. For some mission applications
the gravity gradient torque provides a disturbance which must be countered. However, for
an Earth facing satellite, the gravity gradient provides a simple, low cost means of attitude
stabilization, albeit rather inaccurate. The great advantage is that no fuel is required to
maintain the desired attitude, if a proper stability condition is met. Energy dissipation
is still necessary in order to provide asymptotic convergence onto the prescribed Earth
pointing attitude.

5.3.1

Origin of the gravitygradient torque

The gravity acceleration acting on a mass element dm is given by


~g = GM

~ + ~r
R
~ + ~r ||3
||R

where G is the universal gravity constant, M is the mass of the primary body (the Earth,
~ and ~r are the position vectors of the center of
for an Earth orbiting satellite), while R
mass of the satellite, CM, w.r.t. the centre O of the primary and the position vector of
the mass element dm w.r.t. CM, respectively (Fig. 5.1). The total moment w.r.t. CM is
thus given by
Z
~ =
M
(~r ~g ) dm
B

78

G. Avanzini, Spaceflight Dynamics 5. Environmental torques and other disturbances

2
o
3
e
dm
3
E
3
o

~r

~g

1
o
2
e

CM

~
R
1
E

2
E

1
e
Figure 5.1: Gravity on a finitesize rigid body.


~ + ~r
~r R
~ = GM

dm
M
~ + ~r ||3
||R
B

Since ~r ~r = 0, it is



~ + ~r = ~r R
~
~r R

and the latter expression becomes

~ = GM
M

5.3.2

~
~r R
~ + ~r ||3
||R

dm

Pitch component of the gravitygradient torque

If one is interested only in pitch motion about the axis perpendicular to the orbit plane,
it is possible to obtain a simple expression for the pitch component of the gravity gradient
torque.
~ can be manipulated as follows:
The denominator of the expression of M
h
 
i3/2 h
i3/2
~ + ~r ||3 = R
~ + ~r R
~ + ~r
~ + r2
||R
= R2 + 2~r R

~ respectively. The latter expression can be


where r and R are the magnitude of ~r and R,
rewritten as


2
2
r
3
3
~ + ~r|| = R
~ +
||R
1 + 2 ~r R
3/2
R
R2

Being r R, the third term inside the parentheses can be neglected as it is (r/R)2 1,
and the second one is a first order perturbation compared to 1. But we can expand a
term like (1 + x)n in the neighbourhood of x = 0 as
(1 + x)n = 1 + nx + O(x2 )

G. Avanzini, Spaceflight Dynamics 5. Environmental torques and other disturbances

79

..
...
.
.
. CM
e
dm r ~r
.H
Y
H
..@
.

2 ... @
e
RHH
1 HH R
e
.
~
.
HH
HH
I 
@
3@  orbit
HH
direction
H

H
HH
HH}
O

Figure 5.2: Scheme for the evaluation of gravitygradient torque.


2
~
In the present case, given that x = 2(~r R)/R
and n = 3/2, one gets
!
3/2



2
~
~
r
r

R
3
2
3
3
3
~ +
R
1 + 2 ~r R
R
13 2
=R
1 2 (xX + yY )
R
R2
R
R

1 and e
2 , lie in the orbit
Assuming a pure pitch motion, the first two body axes, e
plane (see Fig. 5.2), so that the resulting pitch motion consists of rotations about the
~ and ~r are
3 only. In terms of body frame components, the position vectors R
body axis e
given by
~ = Xe
1 + Y e
2
R
~r = x
e1 + y
e2
so that

~ = (xY yX)
~r R
e3

3 is thus given by
The torque about e
Z



3
GM
M3 = 3
(xY yX) 1 2 (xX + yY ) dm
R
R
B
Developing the integrand one gets


3
3
(xY yX) 1 2 (xX + yY ) = xY yX 2 (xY yX) (xX + yY )
R
R

3
= xY yX 2 x2 XY + xyY 2 yxX 2 y 2XY
R

and

M3

 Z
Z
GM
= 3 Y
xdm X ydm +
R
B


Z
ZB
Z
Z
3
2
2
2
2

XY
x dm + Y
xydm X
xydm XY
y dm
R2
B
B
B
B

But, for the definition of center of mass, it is


Z
Z
xdm =
ydm = 0
B

and if a frame of principal axis of inertia is chosen as the body frame, also
Z
xydm = 0
B

80

G. Avanzini, Spaceflight Dynamics 5. Environmental torques and other disturbances

and the expression of M3 reduces to


M3

Z

Z
GM
2
2
= 3 5 XY
x dm y dm
R
B
B
Z

Z
GM
2
2
2
2
= 3 5 XY
(x + z )dm (y + z )dm
R
B
B
GM
= 3 5 XY [Jy Jx ]
R

With our choice of axes (see Fig. 5.2), it is


~ = R cos 3 e
1 + R sin 3 e
2 = X e
1 + Y e
2
R
and M3 is thus given by
M3 = 3

5.3.3

GM
cos 3 sin 3 (Jy Jx )
R3

Attitude motion

Since we are considering only pitching motion (1 = 2 = 0), we need only the third
Eulers equation, that can be written in the simple form
Jz 3 = 3

GM
cos 3 sin 3 (Jy Jx )
R3

but, for the same reason, it is also 3 = 3 . Moreover, the quantity


2 =

GM
R3

is the orbital angular velocity, and the equation of the pitching motion under the effect
of gravitygradient becomes
Jy Jx
3 + 32 cos 3 sin 3
=0
Jz
There are two equilibrium solutions, for 3 = 0 and 3 = /2, the stability of which
is determined by the sign of the term Jy Jx . For small perturbations about the local
vertical local horizontal reference frame it is sin 3 3 and cos 3 1, and we get the
second order linear equation of motion


J

J
y
x
2
3 + 3
3 = 0
Jz
If Jy > Jx the term 32 (Jy Jx )/Jz is positive and we can write
2 = 32
The equation of motion becomes

Jy Jx
Jz

3 + 2 3 = 0

the solution of which is given by


3 (t) = 1 cos(t) + 2 sin(t)

G. Avanzini, Spaceflight Dynamics 5. Environmental torques and other disturbances

81

@ 6oscillations
e
@
i
@
@
 PP
@ P
@P
P
2
e
@@
Re
1 PPP
PP
PP
PP
PP
PP}
O









C 
CC 

:


BB e
1
C
e
)
C C C
?
C
divergence
CC
WC

2
e
Figure 5.3: Scheme for the evaluation of gravitygradient torque.
that is, the satellite is stable and oscillates in the neighborhood of the condition 3 = 0
with a frequancy equal to
r
Jy Jx
2 = 32
Jz
If on the converse Jy < Jx , it is 32 (Jy Jx )/Jz < 0, letting
2 = 32

Jx Jy
Jz

the equation of motion can be rewritte as


3 2 3 = 0
and its solution becomes
3 (t) = 1 exp(t) + 2 exp(t)
which is clearly unstable, because of the divergent exponential term. Because of such an
instability, the satellite will rotate away from the neighborhood of the initial condition
3 0, and will acquire a periodic motion around the other equilibrium solution, for
3 = /2. In this case. It is left to the reader as an exercise to demonstrate that, if
Jy < Jx , the equilibrium solution for 3 = /2 is stable.
As an example of the resulting motion, Fig. 5.4 shows the behaviour of the Space
Shuttle (I = diag(1.29, 9.68, 10.01)106 kg m2 ) on a low orbit with a period of 1 hour and
35 minutes. The first simulation is performed starting from an Earth pointing attitude
( = 10 deg), that is the nose of the Shuttle faces the Earth, with a perturbation from the
vertical of 10 deg. The resulting motion is confined and the angular velocity is less tham
with a damping
0.015 deg s1 . Internal dissipation (modeled as a term proportional to ,
constant C = 1.5 106 s1 ) damps very slowly the oscillation in the neighbourhood of the
local vertical.
If on the converse, the motion is started near the second equilibrium condition, for
= 90 deg (that is the nose of the Shuttle points the direction of the orbital velocity),
the resulting motion is unstable and large amplitude oscillation will build up. From the
simulation it can be seen how the vehicle will reverse its attitude, and will oscillate around
the equilibrium at = 90 deg, with the nose pointing towards the external space. It

82

G. Avanzini, Spaceflight Dynamics 5. Environmental torques and other disturbances


10
0.015
0.01
d/dt [deg s1]

[deg]

5
0
5

0.005
0
0.005
0.01
0.015

10
0

4
No. of orbits

10

100

0
[deg]

10

100
[deg]

100

0.1

50
d/dt [deg s1]

[deg]

0
50
100
150
200
250

0.05
0
0.05
0.1

4
No. of orbits

200

Figure 5.4: Stable and unstable motion of the Space Shuttle under gravitygradient
torque.
should also be noted how the effect of the nonlinear term in sin cos in the gravity
gradient torque affects the oscillation frequency, when the amplitude is large. In the last
two plots, the stable motion of the first case is reported as a comparison.

Part II
Advanced Orbital Dynamics: Orbit
Control

83

Chapter 6
Basic concepts
6.1

Introduction

Orbital dynamics plays a central role in space system engineering. Two key functions are
provided:
Mission planning (preflight):
Orbital dynamics is used to assess the feasibility of a desired mission objective and is
used to plan the spacecraft trajectory to achieve mission goals. This may be a brun
sequence for a simple geostationary transfer orbit or a more complex interplanetary
mission with multiple gravity assists. The V budget for the mission is obtained
through simulations which then determine the spacecraft propellant mass. This in
turn links in with the engineering design of the spacecraft. Once a trajectory has
been planned (either a complex interplanetary trajectory or a relatively simple Earth
centered orbit) a timeline of events may be determined from the orbital dynamics,
for example, the periods when the spacecraft is in view of a ground station. Orbital
dynamics is also used to determine sensor coverage patterns for remote sensing and
communication satellites.
Mission Operations (inflight):
Along with preflight mission planning, orbital dynamics is used extensively in
spacecraft mission operations. Principally, orbit prediction is used to determine
the spacecraft location and dynamical state at any given time. This is clearly of importance for planning contacts during ground station passes. Also, orbit prediction
is used to determine the schedule of propulsvie burs for orbit manoeuvres and small
orbit trims. Scheduling is also of importance for planning payload actitivities, such
as activating remote sensing cameras for Earth orbit or interplanetary spacecraft.
The orbit (and the sequence of manoeuvres necessary to achieve it) is determined as a
function of predefined mission objectives. The constraints may be several, such as ground
coverage, for a remote sensing satellite, or planet intercept for an interplanetary mission
with gravity assists.
The best orbit, compatible with mission requirements and constraints, is selected and
the sequence of events, from launch to final orbit injection and operational manoeuvres,
is planned. The correct determination of the V budget is of paramount importance
(as an example, the fuel fraction for orbit control of a communication satellite). Also
the requirements for rendezvouz and docking must be considered during the mission
planning phase and will depend on the capabilities of the vehicle of performing accurate
85

86

G. Avanzini, Spaceflight Dynamics 6. Basic concepts

Mission Objectives
?

Orbit (to satisfy mission)


?

V and transfer times


?

Fuel mass fraction


?

@
I
@
B
B

critical review of
 mission requirements


@
I
@
B
B

constraints on
 orbital manoeuvres


@
I
@
B
B

constraints on mass
 (and payload) fraction


@
I
@
B
B

constraints on
 spacecraft design


DESIGN
Figure 6.1: Mission analysis and design flowchart.
orbit manoeuvres. Once the fuel budget is known, the sizing of the launcher is performed
and the launch windows is determined.
But orbit dynamics is also of crucial importance during mission operations, inasmuch
it is at the base of orbit determination algorithm and orbit control techniques. Moreover,
it may be necessary to perform orbit manoeuvres not scheduled in the original mission
plan (satellite relocation), and the impact on mission effectiveness and satellite lifetime
need to be determined. In this framework, there is a strong need of reducing operational
costs by moving to autonomous operations, that means that orbit determination algorithm
will run on-board and will be coupled with highauthority control systems and mission
planning algorithms.
In what follows, a brief review of basic concepts of orbit motion is proposed, inasmuch
as Keplerian orbits provide a reference solution to the simplest possible dynamic model
of Earthorbiting satellites. In this respect, the presence of other bodies (Moon and Sun
in particular) or other phenomena such as the solar wind or the atmospheric drag will be
seen as a perturbations of this reference solution. The effects of these perturbation will
be the subject of the following chapters. First of all, the variational equations for orbit
parameters will be derived, which provide a means for representing the effect of either

G. Avanzini, Spaceflight Dynamics 6. Basic concepts

87

disturbances or control action on the shape of the orbit and its posizion in the inertial
frame. In the last Chapter, some problems of practical interest will be analyzed, such
as the stationkeeping problem of satellites in LowEarth and Geostationary orbits. In
particular, some examples of control techniques and determination of fuelbudget will be
provided.

6.2
6.2.1

Keplerian motion: a review


Birth of Astrodynamics: Keplers Laws

Since the time of Aristotle the motion of planets was thought as a combination of smaller
circles moving on larger ones. The official birth of modern Astrodynamics was in 1609,
when the German mathematician Johannes Kepler published his first two laws of planetary
motion. The third followed in 1619. From Astronomia Nova we read
... sequenti capite, ubi simul etiam demonstrabitur, nullam Planetae relinqui figuram
Orbit, praeterquam perfecte ellipticam; conspirantibus rationibus, a principiis Physicis,
derivatis, cum experientia observationum et hypotheseos vicari hoc capite allegata.
and
Quare ex superioribus, sicut se habet CDE area ad dimidium temporis restitutorii, quod
dicatur nobis 180 gradus: sic CAG, CAH areae ad morarum in CG et CH diuturnitatem.
Itaque CGA area fiet mensura temporis seu anomaliae mediae, quae arcui eccentrici CG
respondet, cum anomalia media tempus metiatur.
The third law was formulater in Harmonice Mundi
Sed res est certissima exactissimaque, quod proportio quae est inter binorum quorumcunque Planetarum tempora periodica, sit praecise sesquialtera proportionis mediarum
distantiarum, id est Orbium ipsorum
In modern terms, the three laws can be formulated as follows:
First Law: The orbit of each planet is an ellipse with the Sun at a focus.
Second Law: The line joining the planet to the Sun sweeps out equal areas in equal
times.
Third Law: The square of the period of a planet is proportional to the cube of its
mean distance from the Sun.
Kepler derived a geometrical and mathematical description of the planets motion
from the accurate observations of his professor, the Danish astronomer Tycho Brahe.
Kepler first described the orbit of a planet as an ellipse; the Sun is at one focus and
there is no object at the other. In the second law he foresees the conservation of angular
momentum: since the distance of a planet from the Sun varies and the area being swept
remains constant, a planet has variable velocity, that is, the planet moves faster when it
falls towards the Sun and is accelerated by the solar gravity, whereas it will be decelerated
on the way back out. The semi-major axis of the ellipse can be regarded as the average
distance between the planet and the Sun, even though it is not the time average, as
more time is spent near the apocenter than near pericenter. Not only the length of the
orbit increases with distance, also the orbital speed decreases, so that the increase of the
sidereal period is more than proportional.

88

G. Avanzini, Spaceflight Dynamics 6. Basic concepts

Keplers first law can be extended to objects moving at greater than escape velocity
(e.g., some comets); they have an open parabolic or hyperbolic orbit rather than a closed
elliptical one. Yet the Sun lies on the focus of the trajectory, inside the bend drawn by
the celestial body. Thus, all of the conic sections are possible orbits. The second law is
also valid for open orbits (since angular momentum is still conserved), but the third law
is inapplicable because the motion is not periodic. Also, Keplers third law needs to be
modified when the mass of the orbiting body is not negligible compared to the mass of the
central body. However, the correction is fairly small in the case of the planets orbiting the
Sun. A more serious limitation of Keplers laws is that they assume a two-body system.
For instance, the Sun-Earth-Moon system is far more complex, and for calculations of the
Moons orbit, Keplers laws are far less accurate than the empirical method invented by
Ptolemy hundreds of years before.
Kepler was able to provide only a description of the planetary motion, but paved the
way to Newton, who first gave the correct explanation fifty years later.

6.2.2

Newtons Laws of Motion

In Book I of his Principia (1687) Newton introduces the three Axiomata sive Leges Motus:
Lex I - Corpus omne perseverare in statu suo quiscendi vel movendi uniformiter in directum, nisi quatenus a viribus impressis cogitur statum illum mutare.
Lex II - Mutationem motus proportionalem esse vi motrici impress, et fieri secundum
lineam rectam qua vis illa imprimitur.
Lex III - Actioni contrariam semper et qualem esse reactionem: sive corporum duorum
actiones in se mutuo semper esse quales et in partes contrarias dirigi.
The laws of motion can be translated into the following English version:
First Law: Every object continues in its state of rest or of uniform motion in a straight
line unless it is compelled to change that state by forces impressed upon it.
Second Law: The rate of change of momentum is proportional to the force impressed
and is in the same direction as that force.
Third Law: To every action there is always opposed an equal reaction..
The first law requires the identification of an inertial system with respect to which it
is possible to define the absolute motion of the object. The second law can be expressed
mathematically as
~ = d~p
F
dt
~ is the resultant of the forces acting on the object and ~p = m~v is its momentum.
where F
For a constant mass m, it is
~ = m~a
F
where ~a = d~v /dt is the acceleration of the mass measured in an inertial reference frame.
A different equation applies to a variable-mass system (as for instance a rocket launcher
where the variation of the mass per unit time, m,
is sizable). In that case the equation of
motion becomes
~ = m~a + m~
F
v
The third law permits to deal with a dynamical problem by using an equilibrium
~ on m
equation. Its scope is however wider: for instance, the presence of the action F
~ on another portion of an ampler system.
implies an action F

G. Avanzini, Spaceflight Dynamics 6. Basic concepts

6.2.3

89

Newtons Law of Universal Gravitation

In the same book Newton enunciated the law of Universal Gravitation by stating that
two bodies, the masses of which areM and m, respectively, attract one another along the
line joining them with a force proportional to the product of their masses and inversely
proportional to the square of the distance between them, that is,
F =G

Mm
r2

where G = 6.673 1011 m3 kg1 s2 is the Universal Gravitational Constant. The law is
easily extended from point masses to bodies with a spherical symmetry, but it is fairly accurate also for body of arbitrary shape, when their distance is several orders of magnitude
larger than their own dimension.

6.2.4

Equation of Motion in the Two-Body Problem

Consider a system composed of two bodies of masses m1 and m2 , m1 > m2 ; the bodies are
spherically symmetric and no forces other than gravitation are present. Assume an inertial
~ 1 and R
~ 2 describe the positions of masses m1 and m2 , respectively.
frame Fi; vectors R
The position of m2 relative to m1 is
~2R
~1
~r = R
so that their mutual distance is r = ||~r|| and the following equations hold for the relative
velocity and acceleration,
~ 2 R
~ 1 ; ~r = R
~ 2 R
~ 1
~r = R
where all the derivatives are taken in a (at least approximately) inertial frame.
The law of Universal Gravitation is used to express the force acting on each mass, so
that the second law of dynamics can be used to describe separately the motion of each
mass,
~2 = G m1 m2 ~r
~1 = G m1 m2 ~r ; m2 R
m1 R
r3
r3
By subtracting the second equation from the first (after an obvious simplification concerning masses) one obtains
m1 + m2
~r = G
~r
r3
In most practical systems one of the masses is several order of magnitude larger than
the other one. The best balanced 2body system in the Solar system is the planet Pluto
with its satellite Caron, which has a mass equal to 1/7 that of the planet. The Earth
Moon system is the second best balanced pair, where the mass of the Moon is only 0.0123
that of the Earth.
As a consequence, it is usually possible to let m1 = M and m2 = m, with M m, so
that the Equation of Relative Motion can be rewritten in the following simple form:

~r = ~r
r
3

where = GM is the primary body gravitational parameter. Some numerical values of


mu are reported in Tab. 6.1
It is worthwhile to remark that

90

G. Avanzini, Spaceflight Dynamics 6. Basic concepts

the relative position ~r is measured in a non-rotating frame which is not rigorously


inertial;
the relative motion is practically independent of the mass of the secondary body,
when m M; a unit mass will be assumed in the following.

6.2.5

Potential Energy

The mechanical work done against the force of gravity to move the secondary body from
position 1 to position 2
Z 2
Z 2

~r d~s =
L =
dr = +
= U2 U1
3
2
r2 r1
1 r
1 r
does not depend on the actual trajectory from 1 to 2, so that one deduces that the
gravitational field is conservative. The work can be expressed as the difference between
the values of the potential energy at point 1 and 2, the dependence of which on the radius
(i.e. the distance of m from the M) can be expressed in the form
U =

+C
r

The value of the arbitrary constant C is conventionally assumed to be zero in Astrodynamics, that is, the maximum value of the potential energy is zero, when the spacecraft
is at infinite distance from the primary body, otherwise it is negative, and equal to

U =
r

6.2.6

Constants of the Motion

By taking the scalar product of the equation of motion with the velocity vector ~v = ~r ,


~r ~r = 3 ~r ~r
r
and writing everything on the lefthand side by using Eq. (6.6) in Appendix, one gets
 

~v ~v + 3 ~r ~r = v v + 3 r r
r
r
that is,



d v2

=0
dt 2
r
During the motion the specific mechanical energy
v2
E=

2
r
Table 6.1: Values of the gravitational parameter .
Sun
Earth
Moon

1.327 1011
3.986 105
4.903 103

km3 s2
km3 s2
km3 s2

G. Avanzini, Spaceflight Dynamics 6. Basic concepts

91

that is, the sum of kinetic and potential energy, is constant, even though it can be continuously transferred from the kinetic to the potential form, and vice versa. This result
is the obvious consequence of the conservative nature of the gravitational force, which is
the only action on the spacecraft in the two-body problem.
The vector product of ~r with the equation of motion gives

d 
d
~r ~r = 0
~r ~r =
(~r ~v ) = 0
dt
dt

where Eq. (6.3) was applied.


Therefore, the angular momentum

~ = ~r ~v
h
~ is normal to the orbital plane, the spacecraft motion remains
is a constant vector. Since h
in the same plane. This result is not surprising: the unique action is radial, no torque
acts on the satellite, and the angular momentum is constant.
Finally, the cross product of the equation of motion with the (constant) angular mo~ gives
mentum vector h




~ = d ~r h
~ + ~r ~r ~r = 0
~ + ~r h
~r h
r3
dt
r3

By using Eqs. (6.6) and (6.8), one gets

i
d  ~  h 
~r h + 3 ~r ~r ~r (~r ~r ) ~r =
dt
r
!

~r
d ~
r
~r h +
~r
=
dt
r2
r
 
d  ~
d ~r
~r h
= 0
dt
dt r

(6.1)

As a consequence, the vector


~e =

~
~v h
~r

is another constant of the spacecraft motion in the two-body problem.

6.2.7

Trajectory Equation

The dot product


~r ~e =

~
~
~r ~v h
~r ~r
~r ~v h
h2 )

=
r =
r

allows one to determine the shape of the orbit. Letting be the angle between the
(constant) vector ~e and the relative position vector ~r , it is ~r ~e = re cos , so that the
shape of the orbit in polar coordinates centered in M is given by
r=

h2 /
1 + e cos

92

G. Avanzini, Spaceflight Dynamics 6. Basic concepts

P
r

F
a
s

Figure 6.2: Building a conic section.


The radius attains its minimum value for = 0 and the constant vector ~e is therefore
directed from the central body to the periapsis.
A conic is the locus of points P such that the ratio of their distance r from a given
point F (focus) to their distance d from a given line a (directrix) is a positive constant e
(eccentricity), r/d = e. Letting s be the distance between the focus and the directrix it
is (see Fig. 6.2.7)
d = s r cos r = e (s r cos )
By collecting r on the r.h.s., one gets

es
1 + e cos
which is the equation of a conic section, written in polar coordinates with the origin at the
focus. The parameter p = se is called the semilatus rectum. It is the distance between
the focus and the point P of the conic section for = /2.
By comparison with the trajectory equation, one deduces that
r=

in the two-body problem the spacecraft moves along a conic section that has the
primary body in its focus; Keplers first law is demonstrated and extended from
ellipses to any type of conics;
the semi-latus rectum p of the trajectory is related to the angular momentum of the
spacecraft (p = h2 /)
the eccentricity of the conic section is the magnitude of ~e, which is named eccentricity
vector.

6.2.8

Relating Energy and Semi-major Axis

One of the constant of the motion, the angular momentum, has been proved to be simply
related to the semi-latus rectum. Also the specific energy E will now be related to a
geometrical parameter of the conics.

G. Avanzini, Spaceflight Dynamics 6. Basic concepts

93

The constant values of the angular momentum and total energy can be evaluated at
any point of the trajectory, in particular at the periapsis, where the spacecraft velocity is
orthogonal to the radius vector and
h = rP vP ; E =

vP2

h2

= 2
2
rP
2rP
rP

If the orbit is a closed ellipse, the pericentre rP and the apocentre rA (i.e. the closest
and farthest distance of m from the focus where M lies) are given, respectively, by rP =
p/(1 + e) and rA = p/(1 e), so that the semimajor axis a of the ellipse is
p
h2
rP + rA
2
a=
=
p = a(1 e ) =
2
1 e2

It is thus possible to write


h2 = a(1 e2 ) and rP = a(1 e).
By substituting these expressions in that of the specific energy at periapsis, one obtains
a(1 e2 )

E= 2

=
2
2a (1 e)
a(1 e)

1+e
1

2a(1 e) a(1 e)

2a

A very simple relationship exists between the specific mechanical energy and the semimajor axis.
The last relationship between geometrical parameter of the conics and the constant of
the motion is obtained by extracting the eccentricity e from the above expression of h2 ,
that is,
s
s
2
h
h2
e= 1
= 1 2E 2 .
a

It should be noted how degenerate conics have zero angular momentum and therefore unit
eccentricity.

6.2.9

Elliptical Orbits

Geometry of an Elliptical Orbit


An ellipse can be defined as the locus of points such that the sum of their distances from
two fixed point, the focii, is constant:
r1 + r2 = 2a.
The maximum width of the ellipse is the major axis, and its length is 2a, while the
maximum width at the centre in the direction perpendicular to the major axis is the
minor axis, 2b. The distance between the focii is 2c.
From the expression of an ellipses equation in polar coordinates
r=

p
1 + e cos

94

G. Avanzini, Spaceflight Dynamics 6. Basic concepts

and defining its geometrical characteristics as in Fig. 6.2.9, it is easy to derive the following
relations:
rP
rA
2a
2c
b2

=
=
=
=
=

p/(1 + e)
p/(1 e)
rP + rA = 2p/(1 e2 )
rA rP = 2pe/(1 e2 ) = 2ae
a2 c2 = a2 (1 e2 ) = ap

As a consequence, the area of the ellipse


A = ab
can be rewritten as
A = a2

p
p
(1 e2 ) = a (ap)

Figure 6.3: Geometry of an ellipse.


Period of an Elliptical Orbit
The differential element of area swept out by the radius vector as it moves through an
angle d is
1
dA
1
1
h
dA = r 2 d
= r 2 = rv = = const
2
dt
2
2
2
thus demonstrating Keplers third law. The constant value of dA/dt is evaluated by
assuming that the radius vector sweeps out the entire area of the ellipse,
dA
ab
=
dt
T
which permits to obtain the period T of an elliptical orbit:
s
s
2
2
2ab
ab
a3
T =
= 2
= 2
.
h
p

G. Avanzini, Spaceflight Dynamics 6. Basic concepts

95

Figure 6.4: Geocentric frame.

6.3
6.3.1

Three-Dimensional Analysis of Motion


Nonrotating Frames

In order to describe a space mission, it is necessary to provide, together with the time
frame for the sequence of events and evolution of the orbit, a suitable reference system
and a corrisponent set of coordinates, that suits the application. The first requirement
for the spatial description of an orbit is a suitable (and at least approximately inertial)
reference frame. A frame centered in the primary body (i.e. the most massive body
in the system), with axes pointing a fixed direction with respect to the far stars (thus
nonrotating) usually does well the job.
1, g
2, g
3 ) centered in the
For Earth orbiting satellites, a geocentric frame FG = (O; g
center of the Earth and based on the equatorial plane can be used (Fig. 6.3.1), the
equatorial plane being perpendicular to the Earths spin axis. The third unit vector,
namely g 3 , is parallel to the z-axes and is therefore defined; its direction is towards the
North. The x-axis is parallel to the equinox line, that is, the line of intersection of the
ecliptic plane (the plane where the Earth orbit around the Sun lies) and the equatorial
plane; its positive direction is from the Earth to the Sun on the first day of spring or the
vernal equinox. The unit vector g 1 points towards the constellation Aries (the ram).
The Earths axis of rotation actually exhibits a slow precession motion and the xaxis
shifts westward with a period of 26000 years (a superimposed oscillation with a period
of 18.6 years derives from the nutation of the Earths axis, which is due to the variable
inclination of the lunar orbit on the ecliptic plane). For computations, the definition of
the geocentric system is based on the direction of the lineofintersection at a specified
date or epoch. In astronomy, an epoch is a moment in time for which celestial coordinates
or orbital elements are specified; the current standard epoch is J2000.0, which is January

96

G. Avanzini, Spaceflight Dynamics 6. Basic concepts

2 and g 2 , are uniquely defined by the


1st, 2000 at 12:00. The remanining unit vectors, E
requirement for both frames to be orthogonal and righthanded.
The spacecraft position can be described by the Cartesian components of its position
vector with respect to the origin of the reference frame; the use of the distance from
the center and two angles is usually preferred. The declination is measured northward
from the xy plane; the right ascension is measured on the fundamental plane, eastward
from the vernal equinox direction.

6.3.2

Classical Orbital Elements

A Keplerian trajectory is uniquely defined by 6 parameters. The initial values for the
integration of the second-order vector equation of motion, i.e. the spacecraft position
and velocity at epoch, could be used, but a different set of parameters, which provides
an immediate description of the trajectory, is preferable. The classical orbital elements
are widely used for this purpose. Only 4 elements are necessary in the twodimensional
problem: three parameters describe size, shape, and direction of the line of apsides; the
fourth is required to pinpoint the spacecraft position along the orbit at a particular time.
The remaining 2 parameters describe the orientation of the orbital plane.
The classical orbital elements are
1. eccentricity e (shape of the orbit);
2. semimajor axis a or, equivalently, semilatus rectum p (size of the orbit);
3. inclination i, that is, the angle between the third unit vector g 3 and the angular
~ (inclination of the orbit plane with respect to the base plane of the
momentum h
nonrotating reference frame);
4. longitude of the ascending node, , that is, the angle in the fundamental plane
measured eastward from the first coordinate axis of the nonrotating frame g 1 to
the ascending node, i.e. the point where the spacecraft crosses the fundamental
plane while moving in the northerly direction (orientation of the orbit plane);
5. argument of periapsis , that is the angle in the orbit plane between the ascending
node and the periapsis, measured in the direction of the spacecraft motion (pericenter direction in the orbit plane);
6. true anomaly 0 at a particular time t0 or epoch (spacecraft position); it is sometime
replaced by time of periapsis passage T .
Some of the above parameters are not defined when either inclination or eccentricity
are zero. Alternate parameters are
longitude of periapsis = + , which is defined when i = 0 and the ascending
node is not defined;
argument of latitude at epoch u0 = +0 , which is defined when the orbit is circular,
e = 0, and the periapsis is not defined;
true longitude at epoch 0 = + + 0 = + 0 = + u0 , which remains defined
when either i = 0 or e = 0.

G. Avanzini, Spaceflight Dynamics 6. Basic concepts

6.3.3

97

Modified Equinoctial Orbital Elements

It should be noted that geostationary orbits are circular and equatorial, which means that
both the eccentricity and the inclination are zero. One possible alternative choice is to use
the modified equinoctial orbital elements. This is a set of orbital elements that are always
univocally defined, for circular, elliptic, and hyperbolic orbits with any inclination, as
the modified equinoctial equations exhibit no singularity for zero eccentricity and orbital
inclinations equal to 0 and 90 degrees.
However, two of the components are singular for an orbital inclination of 180 degrees,
that is, for retrograde orbits, that is, equatorial orbits flown in the direction opposite to
Earths spin motion. As these orbits are seldom used, this singularity does not represent
a serious problem.
Relationship between modified equinoctial and classical orbital elements is given by
the following set of equations:
1. p = a(1 e2 );
2. f = e cos( + );
3. g = e sin( + );
4. h = tan(i/2) cos ;
5. k = tan(i/2) sin ;
6. 0 = + + 0 .
The inverse relations that allow to recover the classical orbital elements from the modified
ones are as follows:
p
semimajor axis:
a=
1 f 2 g2
p
eccentricity:
e = f 2 + g2


inclination:
i = 2 tan1
h2 + k 2


= tan1 2 h2 + k 2 , 1 h2 k 2
argument of periapsis:

right ascension of the ascending node:


true anomaly at epoch:
argument of latitude at epoch:

= tan1 (g/f ) tan1 (k/h)


= tan1 (gh f k, f h + gk)
= tan1 (k, h)

0 = 0

u0 = + 0
= tan1 (h sin 0 k cos 0 , h cos 0 + k sin 0 )
In the above equations the expression tan1 (a, b) indicates the calculation of the four
quadrant inverse tangent.

6.3.4

Determining the orbital elements

The orbital elements are easily found starting from the knowledge of the Cartesian components of position and velocity vectors ~r and ~v at a particular time t0 in either nonrating

98

G. Avanzini, Spaceflight Dynamics 6. Basic concepts

reference frame defined in Section 6.3.1. One preliminarily computes the components of
the constant vectors
~
~ = ~r v and ~e = ~v h ~r
h

r
and therefore the unit vectors
~
~e
~g 3 ~k
= h ; i = ~r ; n
=
1 =
; p
k
h
r
e
||~g3 ~k||
These vectors, centered in the center of mass of the primary body, define respectively the
directions normal to the orbit plane, and towards the spacecraft, the ascending node, and
the pericenter, respectively.
The orbital elements are thus given by
1.

p = h2 /

2.

e = ||
e||

3.
4.
5.
6.

g 3 = k3
cos i = k

g
1 = n3
cos = n

( > , if n2 < 0)

1
cos 0 = i p

(0 > , if ~r ~v < 0)

p
1
cos = n

( > , if e3 < 0)

In a similar way one evaluates the alternate parameters

cos u0 = i n
(u0 > , if i3 < 0)

1 g
1
cos = p
1
cos 0 = i g

(P i > , if e2 < 0)

(0 > , if i2 < 0)

The last two equations hold only for zero orbit inclination (i = 0).

6.3.5

Orbit propagation: the ideal case

After the orbital elements have been obtained from the knowledge of ~r and ~v at a specified
time, the problem of updating the spacecraft position and velocity is solved in the perifocal
reference frame using the closedform solution of the equation of motion, if the effects
of perturbations is neglected. This means that, provided that the orbit elements remain
(approximately/sufficiently) constant, the satellite will keep on orbiting the Earth along a
known orbit described by the set of 6 parameters chosen. The ~r and ~v components in the
geocentric-equatorial (or heliocentric-ecliptic) frame can be obtained using a coordinate
transformation.

Appendix: Vector Operations


The result of some vector operations is given in this appendix.
~a ~a = a2
~a ~a = 0
~a ~b = ~b ~a
~a ~b ~c = ~a ~b ~c

(6.2)
(6.3)
(6.4)
(6.5)

G. Avanzini, Spaceflight Dynamics 6. Basic concepts

99

Equation (6.5) is intuitively proved by considering that either side represents the volume
of the parallelepiped the sides of which are given by the vectors ~a, ~b, and ~c. In a more
rigorous way, the scalar triple product of three vectors

a1 a2 a3
~a ~b ~c = det b1 b2 b3
c1 c2 c3

is a pseudoscalar and would reverse sign under inversion of two rows. Therefore
~a ~b ~c = ~b ~c ~a = ~c ~a ~b

By equating the first and the last expression, and remembering that the dot product
commutes, one gets the result.
d
d 2
~a ~a = aa
(~a ~a) =
a
dt
dt

(6.6)

Equation (6.6) is very important in orbital mechanics and rocket propulsion: when a vector
elementary increment d~a is considered, its component parallel to ~a increases the vector
magnitude, whereas the perpendicular component just rotates ~a, keeping its magnitude
constant.
As for the vector triple product, the following results hold:




~
~
~
~a b ~c = (~a ~c) b b ~c ~a
(6.7)




~a ~b ~c = ~b ~c ~a (~a ~c) ~b
(6.8)

100

G. Avanzini, Spaceflight Dynamics 6. Basic concepts

Chapter 7
Orbit Trim Manoeuvres
All spacecraft are subject to perturbations which induce a drift in the spacecraft orbital
elements from the nominal mission orbit. As an example, air drag causes a decrease
in semi-major axis, altitude errors cause an alongtrack drift in the spacecraft ground
track and lunar perturbations cause a drift in the inclination of geostationary satellites.
These perturbations must be compensated for by periodic trim manoeuvres to return
the spacecraft to its nominal orbit. The effect of small propulsive burns on the spacecraft
orbital elements can be determined through the use of variational equations. The required
action can then be implemented with impulse thruster firings or continuous acceleration,
by electric propulsion. In particular
cold gas jets are characterised by a low specific impulse Isp , but are very simple
and cheap;
monopropellant thrusters (usually hydrazine reacting on a heated catalyst bed)
provide a good Isp but are expensive;
bi-propellant thrusters provides a higher Isp but are more complex;
electric propulsion is characterised by a very high Isp in a very low force range
(few mN), providing an extended useful life.

7.1

Variational Equations

Consider an orbit frame FO , defined by the unit vectors


r along the outward radial direction (that is from the spacecraft centre of mass
CM towards the outer space, along the line that connects the Earths centre to the
spacecraft CM);
normal to the orbit plane, in the same direction as the satellite angular momentum
n
~r (m~v );
t in the transverse direction (that is perpendicular to r , in the orbit plane and
approximately in the same directions as the orbital velocity vector1 ).
1

To be more precise, the orientation of t is chosen so that t ~v > 0.

101

102

G. Avanzini, Spaceflight Dynamics 7. Orbit Trim Manoeuvres

An external propulsive force can be represented in the orbit frame as

F~ = FR r + FT t + FN n
Also, in this frame, it is
~r = r
r
~v = r
r + r t
The work done (per unit mass) by the external force over an element ds along the
path is
1~ ~
dE = F
ds
m
so that
dE
dt

1~
F ~v
m
1
=
(rF
R + r F
T)
m
=

But for the two body problem, the specific energy of the orbiting body is
E =

2a

and therefore it is also

dE
da
= 2
dt
a dt
Comparing the two expressions of the rate of change of the energy E one gets
da
2a2
=
(rF
R + r F
T)
dt
m

The values of r and can be obtained from the orbit equations for the two body
problem, in which case it is
a(1 e2 )
r=
1 + e cos
Using the chain rule, it is
dr
dr d
r =
=
dt
d dt
so that
dr
a(1 e2 )
d
=
(e sin )
2
dt
(1 + e cos )
dt
e sin
d
= r
(1 + e cos ) dt
Also, from the two body problem, the angular momentum vector magnitude h is given
by
h = r 2 =
and, as a consequence, it is
=

a(1 e2 )

a(1 e2 )
r2

G. Avanzini, Spaceflight Dynamics 7. Orbit Trim Manoeuvres

Considering the mean motion


n=

103

a3

can be rewritten as

na2
1 e2
r2
Substituting the values found for r ad in the expression of the rate of variation of
the semimajor axis, one gets


2a2 na2
F
F
da
re
sin

R
T
=
1 e2
+r
dt
r2
1 + e cos m
m
=

Using the orbit equation,


a
1 + e cos = (1 e2 )
r
the last formulation of a can be further simpliied,

da
2e sin FR 2a 1 e2 FT
=
+
dt
nr
m
n 1 e2 m
It can be noted how the effect of a radial force component FR over the semi-major axis
variation depends on the anomaly , while a force component FT along the transverse
direction has an effect which is independent of the satellite position along the orbit.
In a similar way it is possible to demonstrate that the variational equation for the
eccentricity e, the orbit inclination i and the right ascension are given by
de
=
dt

1 e2 sin FR
+
na
m



1 e2 a2 (1 e2 )
FT
r
2
na e
r
m

di
r cos( + ) FN

=
dt
na2 1 e2 m
d
r sin( + ) FN

=
dt
na2 sin i 1 e2 m
It should be noted how only an outofplane force component can vary the orbit plane
(that is its inclination and/or the ascending node).

7.2

Perigee Raise Manoeuvre

A transverse impulse burn at apogee can be used to raise the perigee, in order to compensate for the effects of air drag. At apogee it is
= ; rap = a(1 + e)

104

G. Avanzini, Spaceflight Dynamics 7. Orbit Trim Manoeuvres

From the variational equation calculated at apogee, with FR = 0, one gets


r
da
2 1 e FT
=
dt
n 1+e m
2
de
FT
=
1 e2
dt
na
m
Assuming that the thruster burn is a pulse of small duration t, the velocity increment
is
FT
VT =
t
m
so that the change of the semimajor axis and orbit eccentricity due to a tangential burn
at perigee is
r
2 1e
a =
VT
n 1+e
e =

2
1 e2 VT
na

Remembering that the perigee radius is


rpg = a(1 e)
the change in perigee radius is
rpg
rpg
a +
e
a
r e
2 1e
2
= (1 e)
VT a
1 e2 VT
n 1+e
na

rpg =

so that
rpg

4
=
n

1e
VT
1+e

With a similar method it is possible to demonstrate that


rap = 0
The change in orbit period
T = 2

r

a3

due to the manoeuvre is

that is

T
1
2
T =
a = 2 p
(3a2 )
3
a
n
2 a /
a2
T = 6

1e
VT
1+e

1e
VT
1+e

G. Avanzini, Spaceflight Dynamics 7. Orbit Trim Manoeuvres

7.3

105

Plane Changes

From the variational equation of orbit inclination


di
r cos( + ) FN

=
dt
na2 1 e2 m
assuming again a pulse of short duration, it is
i =

r cos( + )

VN
na2 1 e2

From this relation, it is evident that, for a given velocity increment VN the variation i
of the orbit inclination is maximised is + = 0, that is the burn has to be applied at
the ascending node, when = .

7.4

Finite Burn Manoeuvres

When lowthrust propulsion is employed (such as electrical thrusters, with a very high Isp
but a thrust in the range of some mN) a burn of finite duration needs to be considered.
In this case it is necessary to perform an explicit integration of the variational equation
rather then simply assuming a finite increment proportional to the rate of variation of the
considered orbit element. Considering a perigee raise manoeuvre as a first example, it is
possible to determine the effects on a and e of a finite burn, for an arclength /2 on
each side of the apogee. Changing the independent variable from time to true anomaly,
one gets
da
da d
=
dt
d dt
so that, assuming a transverse force only, the following equation is obtained:
FT
2a
da na2
1 e2
=
1 e2
nr
m
d r 2
Rearranging, it is
da
2r FT
=
d
n2 a m
2(1 e2 ) FT
=
n2 (1 e cos ) m
For a small trim burn it is possible to neglect the variation of the orbit elements during
the burn, so the total variation of the semimajor axis a is given by
Z off
2
1
2 FT
a = 2 (1 e )
d
n
m on 1 + e cos

where

; off = +
2
2
are the position along the orbit at which the thruster are switched on and off, respectively.
on =

106

G. Avanzini, Spaceflight Dynamics 7. Orbit Trim Manoeuvres

It is possible to explicitly solve the integral


Z off
1
d
on 1 + e cos
using the goniometric transformation
cos =

1 tan2 2
1 + tan2 2

and performing the change of variable


x = tan
such that
The result of the integration is
Z

off

on

and



1 + tan2
d = 2dx
2



+
2
1
2
1e

d =
arctan
tan
1 + e cos
2
1 e2
1 e2
2


+
2
FT
4
1e

2
tan
a = 2 1 e
arctan
n
m
2
1 e2
2

With a similar procedure it is possible to obtain the expression for the increments e
and rpg of the eccentricity and the perigee radius, respectively.
In a similar fashion, it is also possible to analyze the effects of lowthrust propulsion
systems on orbit plane position (represented by and i). As an example, when the
variational equation for orbit inclination is dealt with, one can write
di
r cos( + ) FN
di na2

=
=
1 e2
dt
d r 2
na2 1 e2 m
(that is, the direction normal to the orbit plane) does not
As thruster action along n
change the orbit shape (namely its semimajor axis and eccentricity) it is possible to
integrate the above equation with respect to the true anomaly and evaluate the variation
i induced by a normal acceleration FN /m along an orbit arc from on to off .

Chapter 8
Orbit Control
As stated in the previous chapter, it is possible to vary spacecraft orbit parameters by
propulsive burns. For this reason, the equation that rule orbit shape (a and e) and orbit
plane (i and ) were derived, where orbit parameters increment were determined as a
function of thrust pulse in the orbit reference frame FO .
Such maneuvres may be necessary as a consequence of the effects of external perturbations, which otherwise would result in a spacecraft drifting away from its nominal mission
orbit. It is therefore important that the orbit be controlled through periodic burns to
compensate for these perturbations.
Longitude drift of a geostatonary satellite will lead to the satellite moving out of the
field of view of fixed ground antennae. Similarly, orbit decay will lead to intrack position
errors as the orbital period decrease and the satellite advances ahead of its nominal
position.
At present most orbit control algorithms require a human operator to plan and up
link command sequences to arm the propulsion system and execute the required burns.
However, there is move towards onboard orbit control where the spacecraft uses GPS
satellite navigation to determine its orbital elements and plans its own manoeuvres using
onboard software. This leads directly to significant operational cost reductions, at the
expenses of a much higher cost of the control system, with an equipmenst - hardware and
software - harder to be certified.

8.1

Air Drag Control

For circular Low Earth Orbit (LEO) air drag causes a decrease in orbit radius. It is
possible to use a twoimpulse Hohmann transfer to manoeuvre the spacecraft back to its
nominal orbit.
Example - An Earth observation mission: Orbiting Volcano Observatory (OVO)
Parameter
orbit inclination
altitude
period (approx.)
orbit per day
orbit per week

70 deg
518 km
94 min
15
105

The orbit is specified for mission payload


effectiveness, and there is an exact ground
track repeat requirement.
It is necessary to evaluate the station
keeping fuel budget for maintaining the nominal orbit.

The air drag is quite significant and as orbit decays, the orbit period shortens and the
107

108
a [m]

G. Avanzini, Spaceflight Dynamics 8. Orbit Control


6

518 000 @

517 860

@
@
@
@
@
@
@
@
@
@
@
@
@
@
@
@
6
R
@
@
@
@
@
R
@
@
R
@
@
6
R
@
@
@
@
@
6
@
@
@
@
@
@
@
@
@
@
@
@

dead
band
?
-

4 t [weeks]

exact repeat cycle is lost. The daily altitude loss is approximately 20 m. The resulting
change in orbit period is
r
a
T
T =
a = 3
a
a

which, for a nominal semimajor axis a = 6371 + 518 km and a variation a = 20 m,


gives a variation of the period T = 0.0248 s per day.
The Earth rotates with an angular velocity E of
E =

2
= 7.27 105rad s1
24h

An error of T in orbit period result in a ground track drift of


s = RE E T = 11.5m per day
Over a period of 1 month the ground track drift will be approximately of 320 m, enough
to miss targets with high resolution cameras and narrow field of view. So there is an
evident need to regularly raise the orbit to null the ground track drift.
A 7 day stationkeeping cycle must be designed to raise the orbit of a = 140 m,
using a twopulse Hohmann transfer. Remembering that, for a Hohmann transfer, it is
2
a =
n

1e
V
1+e

the V necessary for each weekly correction is


1
V =
2

a
a3

that, for the case under consideration, is equivalent to V = 0.077 m/s.


Over one year, the V budget is V = 0.077 52 = 4.0 m/s and thte total V
budget for an expected mission life of 3 years is approximately 12 m/s.
At this point, it is necessary to design a spacecraft with enough propellant for a 3
years mission, with a 20% contingency.

G. Avanzini, Spaceflight Dynamics 8. Orbit Control

109

rpg 6
rpg0 @

rpg0 rpg

@
@
@
@
@
@
@
@
@
@
@
@
R
@
@
@
@
R
@
@
R
@
6
@
@
@
6
@
@
@
@
@
@
@
@
@

dead
band
?
-

Example - Perigee control


It is possible to use apogee burn to raise perigee, after air drag decay. For small perturbations, it is possible to assume a linear decrease of the perigee radius with time, that
is
rpg = rpg0 t
where the decay rate can be obtained from flight data of previous satellites or from
simulation. A deadband rpg must be defined, that is the tolerance within which it is
necessary to maintain the satellite perigee radius with respect to the nominal value. If
rpg is within the deadband, no action must be taken, but if rpg crosses the boundary, a
stationkeeping manoeuvre must be performed. In the present case, the manoeuvre is an
apogee tangential burn at the subsequent apogee pass, in order to raise the perigee at the
desired value.
The time between burns is approximately
rpg

t =

The fuel budget for the perigee raise manoeuvre is obtained from the relation
n
V =
4

1+e
rpg
1e

and the station keeping budget (that is the required V per unit time) is
S=
so that
n
S=
4

V
t
r

1+e
1e

It should be noted how the stationkeeping budget is independent of the deadband


amplitude rpg , which is defined by mission objectives, such as payload viewing requirements or similar. Of course, if the tolerance is more strict, more manoeuvres must be
scheduled and, although this does not affect the fuel budget, it increase the mission costs
because of higher workload for the ground staff.

110

8.2

G. Avanzini, Spaceflight Dynamics 8. Orbit Control

Geostationary Orbit Control

Also for geostationary orbits (GEO) the satellite may require accurate stationkeeping, in
order to remain in the field of view of fixed ground antennae for direct broadcasting. After
the launch and the early operation phase (LEOP), during which the satellite is deployed
and injected into its final orbit, the routine operation phase begins. Again, a deadband
must be specified that satisfy accuracy requirements avoiding rapid thruster firings. Of
course, the smaller the deadband, the more accurate is the control, at the expenses of a
greater frequency of manoeuvring.
Also in this case, stationkeeping is usually under ground control, although the trend
is to move to autonomous operations, to reduce operational costs.
There are two types of station keeping modes:
NorthSouth station keeping, required to compensate the effects of lunar and
solar perturbations on orbit inclination, with a significant cost in terms of V for
control;
East-West station keeping, mainly required to compensate the effects of tesseral
harmonics of the Earths gravitational field, induced by the ellipticity of the equator
(which is not a circle).
The location of the satellite on the geostationary ring can be defined in terms of its
longitude, relative to the Greenwhich meridian ( = 0). The rate of change of longitude
is
r

=
E
a3
where E = 2/24 h. For a geostationary orbit, the condition = 0 must be satisfied, so
that it is a =42 164.5 km.
The equatorial crosssection of the Earth is not perfectly circular. Its ellipticity induce
which, for small longitude deadband, can be considered
a transverse acceleration ,
constant. Under this assumption, the variation of satellite longitude on the geostationary
ring is

0 ) (t t0 ) + (t t0 )2
(t) = (t0 ) + (t
2
Given a longitude deadband about the nominal position m , the station keeping
cycle begins at the time t0 and ends at t0 + T . The initial longitude drift rate 0 must be
such that the longitude drift is reversed at the bottom of the deadband, for t = t0 + T /2,
so as to fully exploit the deadband amplitude.
> 0, it is necessary to start the
Assuming a positive longitude drift acceleration
stationkeeping cycle on the East boundary
(t0 ) = m +
0 ) < 0, requiring that the drifting motion is reversed
with a negative initial drift rate (t
at the West bound
(t0 + T /2) = m
where
0 + T /2) = 0
(t

G. Avanzini, Spaceflight Dynamics 8. Orbit Control

111

From the variation of the longitude drift rate

0) +
(t t0 )
(t)
= (t
one gets

+ T /2) = (t
0) +
T = 0
(t
2

so that

0 ) =
T
(t
2
which is negative, as expected. Also, it is
 T 2

T
0) +
(t + T /2) = (t0 ) + (t
= m
2
2 2

Substituting in this equation


0 ) =
T
(t0 ) = m + ; (t
2
one gets

m +
so that

 2  2
T
T
+
= m
2
2 2
 T 2

= 2
2 2

and, finally, the free drift time between burns is given by


r

T =4

Substituting for T into the expression of the initial longitude drift rate, one gets
p

(t0 ) = 2

At the end of the cycle the satellite longitude will return to the top of the deadband,
with
(t0 + T ) = m +
and

0 + T ) = (t
0) = 2
(t

At this point a manoeuvre is required to revert the drift rate of


p

= 4
From the expression of the longitude drift rate
r

=
E
a3

it is easy to obtain the variational equation



3
=
a =
a
2

a
a5

112

G. Avanzini, Spaceflight Dynamics 8. Orbit Control

from which it is evident that a longitude drift rate can be induced by a change a of
the semimajor axis. From the variational equations, a change of a can be obtained by a
transverse burn VT , where
s
a3
a = 2
VT

Therefore

3
= VT
a
and, knowing the required drift change for longitude control, one gets
4 p
VT = a
3

In this way the V required for each cycle of duration T for stationkeeping within a
deadband of amplitude 2 is determined. The stationkeeping budget for mission life
TM with N = TM /T cycles is VM = NV . Multiplication by a safety factor and the
use of the rocket equation allows one to find the required fuel mass.
< 0 can be treated in a similar way, provided
The case of negative drift acceleration
that the stationkeeping limitcycle is started on the West bound with a positive
drift rate, and drift reversal is achieved at the East boundary of the allowed tolerance,
+ .

You might also like