Sie sind auf Seite 1von 30

REACTION ABSTRACT

In this project we try to build a single axis Reaction


Wheel Assembly for a CubeSat.

WHEEL STEVE PRABU


9551174

ASSEMBLY FOR
A CUBESAT
Table of contents

1. Introduction
2. Attitude control methods for spacecraft
2.1. Passive attitude control methods
2.1.1. Gravity gradient
2.1.2. Spin control techniques
2.2. Active attitude control methods
2.2.1. Reaction wheels
2.2.2. Magnetic torquers
2.2.3. Thrusters
3. System requirements for the Reaction wheel Assembly
4. Quantifying the disturbances in the environment
4.1. Solar radiation pressure
4.2. Aerodynamic torque
4.3. Magnetic dipole moment
4.4. Gravity gradient
4.5. Total disturbances
5. Dynamics of spacecraft
6. Investigating variation of inertia matrix for a 1 unit CubeSat
7. Calculating maximum moment of inertia for a 3 unit CubeSat
8. Motor selection
8.1. Working principal of a DC Brushed motor
8.2. Working principal of a DC Brushless motor
8.3. Brushless motor Vs Brushed motor for a Reaction Wheel Assembly
8.4. Parameters of the motor and Electronic Speed Controller chosen
8.5. Limiting constraints of the motor chosen
9. Manoeuvre profiles
9.1. Different manoeuvre profiles
9.2. Manoeuvre profile for minimum system requirements
9.3. Manoeuvre profile using maximum torque of the motor
10. Reaction Wheel sizing
11. Power constraints
12. Summary of the work completed

1
13. Experiments for the future
13.1. Experiment 1: Sensitivity analysis of error due to nonlinear motor
response
13.2. Experiment 2: Dimensionless analysis of a Reaction wheel Assembly
14. References
15. Appendix

List of figures

Figure 1.1. Explorer-1 CubeSat.

Figure 2.2.1.1. A 3 axis reaction wheel assembly.

Figure 2.2.2.1. A CubeSat magnetic torquer.

Figure 4.2.1. Variation of aerodynamic torque with altitude.

Figure 4.3.1. Variation of magnetic torque with altitude.

Figure 4.5.1. Variation of all the disturbances with altitude.

Figure 5.1. Precession due to TN.

Figure 5.2. Depiction of a CubeSat momentum biased along an axis perpendicular to the
orbital plane.

Figure 5.3. Nutation of a gyroscope.

Figure 8.1.1. A DC brushed motor.

Figure 8.2.1. A DC brushless motor.

Figure 9.1. Different manoeuvre profiles to perform a 300 slew manoeuvre in 10s.

Figure 9.3.1. Duration of manoeuvre using maximum available torque.

Figure 9.3.2. Maximum and minimum duration of a slew manoeuvre.

Figure 10.1. Basic dimensions of a reaction wheel.

Figure 13.1.1. Depiction of experiment 1.

2
List of tables

Table 6.1. Dimensions and weigh of different components in a CubeSat.

Table 8.4.1. Motor parameters.

Table 8.4.2. Electronic speed controller parameters.

Table 10.1.1. Basic reaction wheel parameters.

3
1.0 Introduction

The Universe that we live in is expanding at an accelerating rate. This phenomenon


was discovered in 1998 and two projects, namely High-Z Supernova Search Team
and Supernova Cosmology Project started working on gathering more evidence to
support this theory. Eventually, three members from the above two teams were
awarded Nobel Prize for their discovery. These three award-winning discoveries were
based on evidence gathered by the Hubble space telescope. The Hubble space
telescope was able to point in space with very high accuracy because of its attitude
control system.
Attitude is the three-dimensional orientation of a spacecraft with respect to a particular
reference frame. The trajectory of a spacecraft, rocket, missile etc., is mainly governed
by the attitude of the system. Sometimes the sensors present in the spacecraft
measure a vector quantity and hence the sensor might have to point in particular
direction in space in order to measure the parameter. The attitude of a spacecraft must
be controlled in order to point the antennas towards earth for communication purposes
or to expose the solar array to sunlight. The attitude of a spacecraft must also be
stabilised in order to make the spacecraft insensitive to external disturbances.
Although the main purpose of the attitude control system is to orient the spacecraft in
a particular direction in space, it can also be considered as an angular momentum
management system. The attitude control engineer will have to size and decide the
momentum required and the momentum storage capacity depending on the mission
of the spacecraft. The attitude control system of a spacecraft may use one or many
control techniques in order to come up with an optimised solution for the given mission
profile. For example, a spacecraft might be momentum biased when it is in Low Earth
Orbit in order to make it insensitive to external disturbances such as atmospheric drag,
but it need not be momentum biased at certain other altitudes. The required accuracy
of the attitude control system will be determined by the payload and other
‘housekeeping’ subsystems onboard. The frequency of orientation required will also
depend on the time period and eccentricity of the orbit.

4
Today, about 50 years after sputnik was launched, launching a satellite is not just
restricted to space organisations but
even small companies and
universities can launch their own
small satellites. These satellites are
called CubeSats. They have a fixed
dimension and a maximum allowable
weight and hence the attitude control
system of a CubeSat is not just Figure 1.1. Explorer-1 CubeSat.

constrained by power but also by mass and size. In this project, we try to develop a
single axis reaction wheel system that can be used in a 1U, 2U and a 3U CubeSat in
a Low Earth orbit (LEO).

2.0 Attitude Control methods for spacecraft

We can control the attitude of the spacecraft using many methods. Some of which use
inertial properties of spacecraft in order to attain the desired attitude. They are called
passive control methods. Some control methods use power or fuel, and hence they
are called active control methods. Usually, passive control methods have very limited
freedom in terms of repointing manoeuvres.

2.1 Passive attitude control methods

2.1.1 Gravity gradient

A spacecraft usually has a principal axis of rotation that passes through the centre of
gravity and the centre of mass. In this control method, the inertial properties of the
spacecraft makes the principal axis align itself along the local vertical vector. This
vector is called Nadir vector and it always points towards the centre of the Earth. If the
centre of gravity of the spacecraft is at a considerable distance from the centre of
mass, then the spacecraft behaves like a pendulum and oscillates its way back to its
state of least energy i.e., principal axis being parallel to the local vertical. We can use
deployable booms on the spacecraft in order to manipulate the inertia matrix of the
spacecraft to achieve the required attitude.

5
2.1.2 Spin control techniques

In this control method, we leave the spacecraft spinning along a particular axis. This
would create a gyroscopic rigidity along the spinning axis. This would make the
spacecraft insensitive to the external disturbances that create a torque perpendicular
to the spin axis. By doing so we fix the attitude of the spacecraft along the spin axis.
However, this momentum bias would cause cross-coupling while doing a repointing
manoeuvre.

2.2 Active attitude control methods

2.2.1 Reaction wheels

Some space crafts have on board motors with


wheels on them. The reaction torque of these
motors makes the spacecraft rotate. These
wheels are called reaction wheels. This happens
because of conservation of angular momentum.
When these wheels are used along three
perpendicular axes then we can have control over Figure 2.2.1.1. A 3 axis reaction
wheel assembly.
spacecraft’s attitude about the three axes of
rotation.

2.2.2 Magnetic torquers

The Earth’s core creates a magnetic field around


Earth. If the net dipole moment of the spacecraft
is not in line with the local Earth’s magnetic field
then this results in a torque. This torque tries to
align the spacecraft’s dipole moment to be parallel
with the local Earth’s magnetic field. If the Figure 2.2.2.1. A CubeSat magnetic
spacecraft has onboard electromagnets then we torquer.

can manipulate this torque in order to achieve the desired attitude. These
electromagnets are called magnetic torquers.

6
2.2.3 Thrusters

We can also control the spacecraft’s attitude using external thrusters. The thrust vector
from these thrusters should have a moment arm with respect to the centre of mass of
the spacecraft in order to make the spacecraft spin about the desired axis. We can
use cold gas thrusters for attitude control of a small spacecraft. However, we have to
use hydrazine or ion thrusters for larger space crafts.

3.0 System requirements for the Reaction Wheel Assembly

Despite all the available attitude control methods, in this project we shall only consider
a Reaction Wheel Assembly. The given below are self-made system requirements that
would help us constrain few parameters of the design. The system requirements are

1. The reaction wheel assembly should not consume more than 5 W of power.
2. The dimensions of the reaction wheel assembly should not exceed 96mm x
96mm x 15mm.
3. The system should not weigh more than 150 g.
4. The reaction wheel system should be able to recalibrate itself.

4.0 Quantifying the disturbances in the environment

The environment of the spacecraft can produce disturbance torques on the spacecraft.
Theses torques can aggregate with time and make drastic changes to the spacecraft’s
attitude. Some of the main sources of theses disturbances are solar radiation
pressure, atmospheric drag, magnetic dipole moment and gravity gradient. There can
be other disturbances but they are usually very small in magnitude. All these effects
act along their respective centroid i.e., gravitational forces act along the centre of
gravity, solar radiation pressure would act along the centre of pressure and so on. The
reaction wheel assembly must be able to produce more torque than the external
disturbances to have a complete control of the spacecraft’s attitude. Below we
calculate the maximum possible values of these torques.

7
4.1 Solar Radiation pressure

The sunlight consists of small packets of quantised energy called photons. When the
surface of the satellite absorbs these photons, the spacecraft also absorbs their
momentum and the structure experiences a pressure force. When the surface reflects
the photons, the pressure force experienced by the spacecraft is twice as much.
Assuming that the surface has uniform reflectance, the solar radiation torque is given
by

𝜙
𝑇𝑆 = 𝐴𝑆 (1 + 𝑞)(𝑐𝑝𝑠 − 𝑐𝑚)𝑐𝑜𝑠𝜑 … (1)
𝐶

Here TS is the torque because of solar radiation in Nm, 𝜙 is the solar constant (𝜙=1366
W/m2, at distance of 1 astronomical unit from the Sun), c is the speed of light in m/s,
As is the surface area exposed to sunlight in m2, q is the reflectance factor, 𝜑 is the
angle at which sunlight falls on the surface, and cps and cm are the centre of pressure
and the centre of mass of the spacecraft in m. The reflectance factor for a typical
spacecraft is 0.6 (Oland & Sclanbusch, 2009).

The value for solar radiation pressure for a 3U CubeSat is 0.328x10 -8 Nm. This value
fairly remains constant throughout low Earth orbit.

4.2 Aerodynamic torque

In low Earth orbits, the air particles that escape from Earth’s atmosphere exert a force
on spacecraft due to momentum exchange. This effect drastically reduces with altitude
however it is quite significant in low Earth orbit. When this force vector is not aligned
with the centre of mass, it results in a torque. This torque is given as:

1
Ta = 2 ρC𝑑 𝐴𝑟 𝑉 2 (𝑐𝑝𝑎 − cm) … (2)

Here Ta is the torque due to atmospheric drag in Nm, ρ is the density of atmosphere
in Kg/m3, Cd is the coefficient of drag for the spacecraft, Ar is the frontal area of the
spacecraft as seen from the incoming flow (also called as the ram area) in m 2, V is the
velocity of spacecraft in orbit in m/s, cpa and cm are the corresponding centroids for
atmospheric pressure force and centre of mass in m. The coefficient of drag of a
spacecraft varies between 2 and 2.5 (Wertz, et al., 2011). The density and orbital

8
velocity of spacecraft is a function of altitude. In figure 4.2.1, we plot the variation of
aerodynamic torque with altitude. Note that the y-axis is shown in log scale.

0
100 200 300 400 500 600 700 800 900

-2
Aerodynamic torque (Log10 scale)

-4

-6

-8

-10

-12
Atitude in Km

Figure 4.2.1. Variation of aerodynamic torque with altitude.

The maximum possible aerodynamic torque occurs at an altitude of 100 Km and has
a value of 0.005923 Nm.

4.3 Magnetic dipole moment

Earth has a magnetic field around it because of its liquid core. When the residual
magnetic dipole of the spacecraft is not in line with the Earths local magnetic field then
it experiences a magnetic torque. This torque tries to align the spacecraft’s net dipole
moment to be in line with Earth’s magnetic field. The magnitude of this torque is given
by the following equation

𝑀
𝑇𝑚 = 𝐵𝐷 = 𝐷 (𝑅3 𝜆) … (3)

Where Tm is the magnetic torque acting on the spacecraft in Nm, B is Earth’s magnetic
field strength in tesla, D is the spacecraft’s residual dipole moment in A.m2, M is the
magnetic constant multiplied by Earth’s magnetic moment in tesla.m3, R is the
distance from the centre of Earth in m and 𝛌 is a unit less function that varies from 1
at the equator to 2 at the poles (Wertz, et al., 2011). For a large spacecraft the residual

9
dipole moment varies between 0.2-20 A.m2. However it is about 0.1 for a 3U CubeSat
(Garza, 2008). The variation of magnetic torque on spacecraft with altitude is shown
in figure 4.3.1.

0.000006

0.000005
Magnetic torque in Nm

0.000004

0.000003

0.000002

0.000001

0
100 200 300 400 500 600 700 800 900
Altitude in Km

Figure 4.3.1. Variation of magnetic torque with altitude.

The maximum magnetic torque on a spacecraft acts at an altitude of 100 Km and has
a value of 5.6805x10-6 Nm.

4.4 Gravity gradient

When the centre of gravity of the spacecraft is not in line with the centre of mass, with
respect to the local Nadir vector, then the spacecraft experiences gravity gradient
torque. The principal axis of rotation always goes through the centre of mass and the
centre of gravity. The magnitude of this torque increases with the angle between the
principal axis and the Nadir vector. The magnitude of this torque is given by


𝑇𝑔 = 2𝑅 |𝐼𝑍 − 𝐼𝑌 |𝑆𝑖𝑛(2ɵ) … (4)

Here Tg is the torque in Nm, µ is the Earth’s gravitational constant in m 3/s2, ɵ is the
angle between the principal axis and the Nadir vector, R is the distance from the centre
of Earth in m, Iz and IY are the moments of inertia of the spacecraft in Kg.m2. The
maximum possible value of this torque for a 3U CubeSat is 3.3x10-8 Nm (Garza, 2008).

10
4.5 Total disturbances

All the above torques were calculated for different altitudes under the worst case
conditions. The variation of disturbance torque with altitude is shown in figure 4.5.1.

0
100 300 500 700 900
-2
Torque (log10 scale)

-4 SRP
Drag
-6 Gravity gradient
Magnetic torque
-8

-10

-12
Altitude in Km

Figure 4.5.1. Variation of all the disturbances with altitude.

The maximum disturbance occurs at an altitude of 100 Km and has a value of


0.005928 Nm.

5.0 Dynamics of spacecraft


The spacecraft we will be considering in our calculations shall have a non-spinning
structure along with spinning wheels in the inside for dynamic properties such as
gyroscopic rigidity and for providing internal reaction torques for spacecraft slew
manoeuvre. This combination of an inertially fixed structure and internal spinning
wheels enable us to accommodate a large variety of payload into very simple
structures. This allows us to use pencil-shaped space crafts that can fit better in launch
vehicles.

The dynamics of spacecraft attitude obeys the Euler equation for rotational dynamics.
Euler equation expressed in terms of spacecraft’s reference frame is:
Ḣ= 𝑇−𝜔×𝐻 … (5)

11
The above equation is based on conservation of angular momentum (H). From the
above equation, it can be seen that internal torques will not change the total angular
momentum of the system. For example, the forces due to any internal mechanisms or
astronauts moving inside the spacecraft will not change the total angular momentum
of the system. The component of external torque that is in the same direction of
angular momentum will only change the magnitude of momentum. The component of
external torque TN that is at right angles to the angular momentum will change the
direction of angular momentum. This change of angular momentum vector H by torque
component TN is called precession. From figure 5.1, it can be seen that rate of
precession is given by
𝑑𝛹 𝑇𝑁
= … (6)
𝑑𝑡 𝐻

TNdt

Figure 5.1. Precession due to TN .

The total angular momentum of the spacecraft can be written as the sum of the angular
momentum of the spacecraft along with the angular momentum of the spinning wheels
inside.
𝐻 = 𝐼𝜔 + ℎ … (7)
Here I is the inertia matrix of the spacecraft and h is the angular momentum of any
rotating objects inside the spacecraft, such as reaction wheels or momentum wheels.
In matrix form, the above equation can be written as
𝐻𝑋 𝐼𝑋𝑋 −𝐼𝑋𝑌 −𝐼𝑋𝑍 𝜔𝑋 ℎ𝑋
[𝐻𝑌 ] = [−𝐼𝑌𝑋 𝐼𝑌𝑌 −𝐼𝑌𝑍 ] [𝜔𝑌 ] + [ℎ𝑌 ]
𝐻𝑍 −𝐼𝑍𝑋 −𝐼𝑍𝑌 𝐼𝑍𝑍 𝜔𝑍 ℎ𝑍
𝐻𝑋 𝐼𝑋𝑋 𝜔𝑋 − 𝐼𝑋𝑌 𝜔𝑌 − 𝐼𝑍𝑋 𝜔𝑍 + ℎ𝑋
[𝐻𝑌 ] = [ 𝐼𝑌𝑌 𝜔𝑌 − 𝐼𝑌𝑍 𝜔𝑍 − 𝐼𝑋𝑌 𝜔𝑋 + ℎ𝑌 ]
𝐻𝑍 𝐼𝑍𝑍 𝜔𝑍 − 𝐼𝑍𝑋 𝜔𝑋 − 𝐼𝑌𝑍 𝜔𝑌 + ℎ𝑍

Every spacecraft has a set of orthogonal axis of rotation called the principal axis for
which the inertia matrix becomes a diagonal matrix, i.e.

12
𝐼𝑋𝑋 0 0
𝐼𝑃𝑟𝑖𝑛𝑐𝑖𝑝𝑎𝑙 =[ 0 𝐼𝑌𝑌 0]
0 0 𝐼𝑍𝑍

Differentiating equation (7) with respect to time gives


Ḣ = İ𝜔 + 𝐼ώ + ℎ̇ … (8)

Substituting equation (8) in equation (5) gives

İ𝜔 + 𝐼ώ + ℎ̇ = 𝑇 − 𝜔 × 𝐻

𝐼ώ = 𝑇 − ℎ̇ − İ𝜔 − 𝜔 × 𝐻 … (9)

The above equation shows us how the attitude of the spacecraft changes with various
factors that cause it. T represents all the external torques that act on the system. This
may be caused by one or many of the external disturbances such as solar radiation
pressure, aerodynamic forces, the magnetic dipole moment of the spacecraft or
gravity-gradient. The term ℎ̇ accounts for all the changes of angular momentum of all
the internal rotational objects such as momentum wheels and reaction wheels. İ𝜔
shows how the changes in the inertia matrix of the spacecraft, such as by deploying
solar arrays, can affect the attitude dynamics of the spacecraft. The term 𝜔×H is called
the gyroscopic torque.

In equation (9) if we assume that the mass distribution of the spacecraft will not
change, then we can approximate the term İ𝜔 to be zero.

Now equation (9) in matrix form is

ℎ𝑋̇ 𝑇𝑋
[ℎ𝑌̇ ] = [𝑇𝑌 ] +
ℎ𝑍̇ 𝑇𝑍

𝐼𝑋𝑌 𝜔𝑌̇ − 𝐼𝑋𝑋 𝜔𝑋̇ + 𝐼𝑋𝑍 𝜔𝑍̇ − 𝐼𝑍𝑍 𝜔𝑍 𝜔𝑌 + 𝐼𝑍𝑋 𝜔𝑋 𝜔𝑌 + 𝐼𝑌𝑍 𝜔𝑌2 + 𝐼𝑌𝑌 𝜔𝑍 𝜔𝑌 − 𝐼𝑌𝑍 𝜔𝑍2 − 𝐼𝑌𝑍 𝜔𝑋 𝜔𝑍
[𝐼𝑋𝑌 𝜔𝑋̇ − 𝐼𝑌𝑌 𝜔𝑌̇ + 𝐼𝑌𝑍 𝜔𝑍̇ − 𝐼𝑋𝑋 𝜔𝑋 𝜔𝑍 + 𝐼𝑋𝑌 𝜔𝑍 𝜔𝑌 + 𝐼𝑋𝑍 𝜔𝑍2 + 𝐼𝑍𝑍 𝜔𝑋 𝜔𝑍 − 𝐼𝑍𝑋 𝜔𝑋2 − 𝐼𝑍𝑋 𝜔𝑋 𝜔𝑌 ]
𝐼𝑋𝑍 𝜔𝑋̇ − 𝐼𝑍𝑍 𝜔𝑍̇ + 𝐼𝑋𝑌 𝜔𝑌̇ − 𝐼𝑌𝑌 𝜔𝑋 𝜔𝑌 + 𝐼𝑌𝑍 𝜔𝑋 𝜔𝑍 + 𝐼𝑋𝑌 𝜔𝑋2 + 𝐼𝑋𝑋 𝜔𝑋 𝜔𝑌 − 𝐼𝑋𝑌 𝜔𝑌2 − 𝐼𝑋𝑌 𝜔𝑌 𝜔𝑍

… (10)

13
The above equation is a very generic equation. Since in the above equation, motion
about all three axes are related, a reaction torque about a single axis can set the
spacecraft tumbling about all three axes. However, by having a 3 axis control system
we can produce reaction torque in two or more axis which results in the required
repointing or slew manoeuvre without making it tumble. But for a CubeSat to have a 3
axis reaction wheel attitude control system is not ideal since it would occupy a lot of
space and require a lot of power. Hence for a CubeSat we try and make wheels spin
along the principal axis. By doing so the inertia matrix would become a diagonal matrix
and the wheels would not cause a rotation in any of the perpendicular axes. Thus for
a spacecraft that has a diagonal inertia matrix, equation (10) becomes

ℎ𝑋̇ 𝑇𝑋 𝐼𝑋𝑋 𝜔𝑋̇ − 𝐼𝑍𝑍 𝜔𝑍 𝜔𝑌 + 𝐼𝑌𝑌 𝜔𝑍 𝜔𝑌


[ℎ𝑌̇ ] = [𝑇𝑌 ] + [𝐼𝑌𝑌 𝜔𝑌̇ − 𝐼𝑋𝑋 𝜔𝑋 𝜔𝑍 + 𝐼𝑍𝑍 𝜔𝑋 𝜔𝑍 ] … (11)
ℎ𝑍̇ 𝑇𝑍 𝐼𝑍𝑍 𝜔𝑍̇ − 𝐼𝑌𝑌 𝜔𝑋 𝜔𝑌 + 𝐼𝑋𝑋 𝜔𝑋 𝜔𝑌

To provide gyroscopic rigidity, CubeSats are momentum biased along an axis that is
perpendicular to the orbital plane of the spacecraft as shown in figure 5.2.

Axis of momentum
bias.

Figure 5.2. Depiction of a CubeSat momentum biased along an axis


perpendicular to the orbital plane.

When the angular momentum vector is not perfectly in line with the principal axis, then
the spacecraft wobbles about the precession axis. This is called nutation. This is
shown in the figure 5.3.

14
Figure 5.3. Nutation of a gyroscope.

For example if the spacecraft is momentum biased about the z axis then angular
velocity of nutation is given by

𝐻𝑍
𝜔𝑛𝑢𝑡 = … (12)
√𝐼𝑋𝑋 𝐼𝑌𝑌

6.0 Investigating variation of inertia matrix for a 1 unit CubeSat

In all our calculations we assume that the reaction wheel assembly is along the
principal axis of rotation. However this is not always true. Hence we try to analyse how
the inertia matrix varies using solidworks. A basic CubeSat would have transceiver
system, a power supply, 3 axis cold gas system, attitude control system and an on-
board computer. The typical dimensions of these components were approximated
using values from CubeSatshop.com. The dimensions and weight of these
components are tabulated in table 6.1.

Length (mm) Width (mm) Height(mm) Weight(g)


Transceiver 96 90 15 75
Power supply 96 90 7 80
3 axis CGS 96 96 15 300
system
ASC board 96 96 15 200
On-board 96 90 12.4 94
computer
Structure 100 100 113.5 87.2
Table 6.1. Dimensions and weigh of different components in a CubeSat.

15
Model of these components was made in solidworks and assembled in all possible
permutations and combinations. The inertia matrix about the geometric centre for all
these combinations was also calculated. The inertia matrix for extreme mass
distributions along with the average value of the inertia matrix is given below.

0.001893 1.756𝑥10−5 5.85𝑥10−6


𝐼𝑒𝑥𝑡𝑟𝑒𝑚𝑒,1 = [1.756𝑥10−5 0.00146 4.39𝑥10−5 ] Kg.m2
5.85𝑥10−6 4.39𝑥10−5 0.001911

0.001323 −3.51𝑥10−5 5.9𝑥10−6


𝐼𝑒𝑥𝑡𝑟𝑒𝑚𝑒,2 = [−3.51𝑥10−5 0.001454 −3.5𝑥10−5 ] Kg.m2
5.9𝑥10−6 −3.5𝑥10−5 0.00134

0.001551 −3.51𝑥10−6 8.78𝑥10−7


𝐼𝑎𝑣𝑒𝑟𝑎𝑔𝑒 = [−3.51𝑥10−6 0.001456 −1.8𝑥10−6 ] Kg.m2
8.78𝑥10−7 −1.8𝑥10−6 0.001567

The inertia matrix of every combination is tabulated in Appendix 1.

Assumptions made in the inertia matrix calculations is listed below.

 We assume that all the components have homogeneous mass distribution.


 We also assume that the CubeSat does not have any other masses like sensors
or deployable solar panels.
 Since the transceiver, power supply and the on-board computer have similar
weight, we assume that all three of them have the average weight (i.e. 83 g) in
order to reduce the number of combinations possible.

7.0 Calculating maximum moment of inertia for a 3 unit CubeSat

Below is the inertia matrix of a 1U CubeSat measured about the geometrical centre
under extreme mass distribution circumstances.

0.001893 1.756𝑥10−5 5.85𝑥10−6


𝐼𝑒𝑥𝑡𝑟𝑒𝑚𝑒,1 = [1.756𝑥10−5 0.00146 4.39𝑥10−5 ] Kg.m2
5.85𝑥10−6 4.39𝑥10−5 0.001911

From the inertia matrix analysis of 1U CubeSat we get that the maximum moment of
inertia about an axis is 0.001911 Kg.m3. From this we can calculate the maximum
moment of inertia value for a 3U using parallel axis theorem. The parallel axis theorem
is given by;

16
𝐼 = 𝐼𝑐𝑚 + 𝑚𝑑 2 … (13)

Here I is the moment of inertia about the centre of mass of the structure in Kg.m3, m
is the mass of structure in Kg and d is the distance between the old axis of rotation
and the new axis of rotation in m.

From equation (13) we get that the maximum moment of inertia about an axis for a 3U
CubeSat is 0.025969 Kg.m2.

8.0 Motor selection

A CubeSat has a DC power source. Thus we should use a DC motor to run the reaction
wheels. There are two types of DC motor. They are

 Brushed motor
 Brushless motor

The working principal of both the types of motors are explained below.

8.1 Working principal of a DC Brushed motor

A brushed DC motor has permanent magnets


along the outside. These permanent magnets do
not move and hence they are called stator. The
motor has a spinning armature in the inside, and
since it rotates it is called rotor. The armature
inside the motor contains an electromagnet.
When electricity is passed through this
electromagnet it creates a magnetic field around
Figure 8.1.1. A DC brushed motor.
it. These electromagnets get attracted to the
permanent magnets and starts rotating. After the rotor turns by 180 degrees the
brushes change the polarity of the electromagnet in order to keep the rotor spinning.

17
The disadvantages of using a brushed motor are

 Since the brushes are constantly breaking and making connections, they wear
out eventually. They also cause sparks and electrical noise.
 The maximum speed of a brushed motor is limited by the brushes.
 The electromagnet takes time to cool down since it is in the inside of the motor.

8.2 Working principal of a DC Brushless motor

On the contrary to the brushed motor, in a


brushless DC motor the rotor is the permanent
magnet and the stator is the electromagnet.
These electromagnets are alternatively polarised
to make the permanent magnet in the inside
rotate. The brushless motor is controlled by an
electronic speed controller that require a pulse
width modulation input to control the speed of the Figure 8.2.1. A DC brushless motor.

motor. Since the brushless motor uses input duty


cycle to run, we have a greater control over the speed of the motor compared to a
brushed motor. Since a brushless motor does not use brushes, there is much lesser
electrical noise and there is no sparks created. The electromagnets cool down easily
since it is along the outside of the motor. It is also possible to increase the control we
have over the speed of the motor if we use a motor with more number of
electromagnets along with a high resolution electronic speed controller.

8.3 Brushless motor Vs Brushed motor for a Reaction Wheel


Assembly

Since the advent of semiconductors and transistors, brushless motors have improved
a lot compared to brushed motors. The brushless motors have become more complex
and advanced. In this era of electronics a brushless motor enables the user to have
more precise control over its speed. Brushless motors also have a longer life than a
brushed motor since it uses no brushes. Since it does not use brushes it does not

18
suffer from brush friction and hence a brushless motor exhibits a higher linearity
compared to a brushed motor. Hence we will use a brushless motor for our reaction
wheel assembly.

8.4 Parameters of the motor and Electronic Speed Controller chosen

Parameter Unit
Poles 9N12P -
KV 2800 RPM/V
Input Voltage 7.4 V
Weight 12.8 g
Table 8.4.1. Motor parameters.

Parameter unit
Max current 7 Amp
PWM 8~16 KHz
Weight 7 g
Table 8.4.2. Electronic speed controller parameters.

8.5 Limiting constraints of the motor chosen

A brushless motor has a parameter KV that relates the RPM of the unloaded motor to
the voltage applied across the motor.

𝑅𝑃𝑀 = 𝐾𝑉 𝑉 … (14)

The selected brushless motor has a KV value of 2800 RPM/V. The motor also has a
maximum input voltage of 7.4 V. However a typical voltage supply from the power
source in a CubeSat is 3.7 V. Multiplying KV value with 3.7 V gives the maximum
angular velocity of the wheel to be 1085 rad/s. The electronic speed controller used
has a pulse width modulation resolution of ~8-16 Hz. Under the worst case the
resolution of the electronic speed controller would be 8 Hz. Hence the smallest
increment in angular velocity would be the angular velocity that corresponds to one

19
resolution point .i.e., 0.13563 rad/s. Thus the smallest possible increment in angular
velocity by the reaction wheel system would be 0.13563 rad/s.

9.0 Manoeuvre profiles

For the reaction wheel assembly one of the system requirements mentioned earlier is
that it should be able to make a 3U CubeSat perform a ±30 0 slew manoeuvre within
10 seconds.

Now this manoeuvre can be done in many ways. The angular velocity of the wheel in
some of the possible manoeuvre profiles are shown in figure 9.1.

0.6
Angular velocity of the wheel in rad/s

0.5

0.4

0.3 Slow ramp


Impulsive
0.2
Composite

0.1

0
0 2 4 6 8 10 12
time in s

Figure 9.1. Different manoeuvre profiles to perform a 300 slew manoeuvre


in 10s.

One extreme of the different possibilities is shown by the red curve. In this we
accelerate the wheel as fast as possible with an impulsive torque and then allow the
wheel to coast at a constant speed for few seconds before braking the wheel to a stop.
The other extreme is to accelerate the wheel slowly for 5 seconds and decelerate the
wheel for the next 5 seconds as shown by the blue curve. However the area under all
these curves are equal since it gives the angle of the slew manoeuvre. Between these
two extremes there are infinite number of possibilities by which we can manoeuvre the
spacecraft. However we will perform all our manoeuvre calculations for the manoeuvre
that follows the blue curve. This is because we do not want any tumbling because of

20
the impulsive jerk and also because of the fact that motor response might not be linear
at high accelerations. Hence a slow acceleration would allow us to reduce the error
due to non-linear response of the motor. Below we try to calculate the boundary
conditions for the manoeuvre profile. The minimum condition is derived from the
system requirement and the maximum condition is derived from the physical limitation
of the motor.

9.2 Manoeuvre profile for minimum system requirements

During the manoeuvre we can say that the torque acting on the spacecraft is equal to
the reaction torque of the reaction wheel. The torque acting on the spacecraft is given
by

⃗ = 𝐼𝛼
𝑇 … (15)

Where I is the maximum possible moment of inertia for a 3U CubeSat and α is the
angular acceleration of the spacecraft. From kinematics we can tell that

1
𝛽 = 𝛽0 + 𝜔0 (𝑡 − 𝑡0 ) + 2 𝛼(𝑡 − 𝑡0 )2 … (16)

Here β is the angle of the spacecraft in radians, 𝜔 is the angular velocity of spacecraft
in rad/s, α is the angular acceleration of the spacecraft in rad/s2 and t is the time in s.
Substituting equation (16) in equation (15), we get

2𝐼
𝑇 = (𝑡−𝑡 2
[𝛽 − 𝛽0 − 𝜔0 (𝑡 − 𝑡0 )] … (17)
0)

Since the reorientation manoeuvre is from rest to rest and the duration of angular
acceleration is t/2, equation (17) implies

4𝛽𝐼
𝑇= … (18)
𝑡2

Substituting the values in equation (18) gives us T=0.000543893 Nm. Since the torque
acting on spacecraft equals the reaction torque of the wheel we get

0.000543893 = 𝐼𝑤ℎ𝑒𝑒𝑙 𝛼𝑤ℎ𝑒𝑒𝑙 … (19)

Using the minimum torque equation (18) implies

𝑡 2 = 3.333336178 ɵ … (20)

21
9.3 Manoeuvre profile using maximum torque of the motor

A brushless motor has another constant called as the Kt torque constant. This constant
relates the torque created by the motor to the current driven by the motor. The equation
is given below;

𝑇 = 𝐾𝑡 𝐼 … (21)

In the above equation T is the torque of the motor in Nm, Kt is the torque constant and
I is the current driven by the motor in Amperes. However the torque constant of a
motor relates to the KV value of a motor by the following relationship

1
𝐾𝑡 = 𝐾 … (22)
𝑉

Using equation (22) we get the torque constant to be equal to 0.000357 N.A -1. Since
the selected brushless motor is rated 6 Amp, equation (21) implies that the maximum
torque (Tmax) produced by the motor is 0.0021429Nm. Using the value of Tmax and Imax
in equation (18) gives us the relationship between the time of manoeuvre and the
angle by which the spacecraft turns.

𝑡 2 = 0.8460396 ɵ … (23)

The plot of equation (23) is given in figure 9.3.1.

20
18
16
Time for manoeuvre (s)

14
12
10
8
6
4
2
0
0 50 100 150 200 250 300 350
Manoeuvre angle in degrees

Figure 9.3.1. Duration of manoeuvre using maximum available torque.

22
From the equation (23) we get that the shortest possible duration for a 300 manoeuvre
is 5.04 seconds. But the system requirement states that the maximum duration for a
300 manoeuvre shall not exceed 10s. The above two conditions gives us the boundary
condition for the torque of the reaction wheel system i.e., 0.0005439 Nm ≤ T=I wheel
𝜔̇ wheel ≤ 0.0021429 Nm. The minimum and the maximum possible duration for a slew
manoeuvre is shown in figure 9.3.2.

40

35
Time for manoeuvre (s)

30

25

20 Maximum available
torque
15
Minimum alowable
10 torque

0
0 100 200 300
Manoeuvre angle in degrees

Figure 9.3.2. Maximum and minimum duration of a slew manoeuvre.

The manoeuvre time profile of the reaction wheel assembly should lie between the two
lines in the above figure.

10.0 Reaction Wheel sizing

The basic dimensions of a reaction wheel is shown in figure 10.1.

Figure 10.1. Basic dimensions of a reaction wheel.

23
The moment of inertia for the above wheel is given by
𝜋 4 4 4
𝐼𝑟𝑤 = 𝜌 2 [ℎ𝑟𝑖𝑛𝑔 (𝑟𝑟𝑤 − 𝑟𝑑𝑖𝑠𝑘 ) + ℎ𝑑𝑖𝑠𝑘 𝑟𝑑𝑖𝑠𝑘 ] … (24)

The design parameters are tabulated in table 10.1.1.

Parameter Value Unit


Material Aluminium -
Density 2700 Kg/m3
hring 0.005 m
hdisk 0.001 m
rrw 0..04 m
rdisk 0.03 m
Ire 4.40095E-05 Kg.m2
mass 0.0405108 Kg
Table 10.1.1. Basic reaction wheel parameters.

For the designed wheel the minimum required angular acceleration to do a 300
manoeuvre within 10s is 12.359 rad/s2.

11.0 Power constraints

An approximate value of the power consumed by the reaction wheel assembly can be
calculated by dividing the kinetic energy of the wheels by the efficiency of the brushless
motor.
1
𝐼𝑤ℎ𝑒𝑒𝑙 𝜔 2
𝑃𝑜𝑤𝑒𝑟 = 2 … (25)
𝜂

Here Power is in W, Iwheel is in Kg.m2, 𝜔 is the angular velocity of the wheel in rad/s
and η is the efficiency of the brushless motor. The typical efficiency of a brushless
motor is between 0.8 and 0.9. Thus for the designed wheel maximum allowable
angular velocity for a 5 W power consumption is 426 rad/s.

24
12.0 Summary of the work completed

So far we have calculated the maximum possible moment of inertia for a 3 Unit
CubeSat under extreme mass distributions, external disturbances in low Earth orbit,
dimensions and material of the reaction wheel and decided the motor and electronic
speed controller. We have also calculated the minimum possible increment in angular
velocity of the wheel that is constrained by the resolution of ESC and the KV value of
motor, the manoeuvre duration profile for minimum and maximum motor torque and
the maximum allowable angular velocity that is constrained by power consumption.

13.0 Experiments for the future

At this point we have two main issues:

First is that we do not yet have an optimised acceleration for the wheel. One of the
ways to decide the angular acceleration is by making sure the motor response is linear
with the acceleration demanded by the user.

Second is that later if we decide on changing the moment of inertia of the wheel or
change the motor or use the wheel for a different spacecraft, we may have to go
through the whole process again for the new reaction wheel system. This can avoided
if we develop a dimension less relation for the reaction wheel system that is
independent of motor, wheel and spacecraft.

To deal with the above issues we shall conduct the following experiments in order to
come up with a solution.

13.1 Experiment 1: Sensitivity analysis of error due to nonlinear


motor response

The response time of a motor is not linear for all angular accelerations of the wheel.
This error increases with very high angular accelerations. In order to analyse motor
response with very high resolution, we use the wheel as a photo gate. The wheel has
small holes along the perimeter that can be used as photo gate for the IR LED and IR
sensor. The IR sensor will be connected to a data acquisition board that sends the

25
data to a computer in order to plot the acceleration of the wheel in real time. A depiction
of the experiment is shown in figure 13.1.1.

IR sensor
IR LED

DAQ

Motor

Computer

Reaction wheel

Battery

Figure 13.1.1. Depiction of experiment 1.

However the wheel should have the same moment of inertia as the reaction wheel
sized in table 10.1.1. This is done in order to make sure that the load on the motor is
the same. We can increase the resolution of the experiment by increasing the number
of holes.

Sources of error

 The holes might not be uniformly spaced.


 The wheel might be spinning at a speed faster than the data acquisition system.
 There might be external IR noise.

13.2 Experiment 2: Dimensionless analysis of a Reaction wheel


Assembly

In this analysis we try to come up with a function that is universal for a reaction wheel
system that is independent of the motor, size of the wheel and the moment of inertia
of the spacecraft. Let us say that the duration (t) of a manoeuvre depends on the input
26
voltage (Vin) of the motor, KV value of the motor, the rate of change of angular
momentum (ℎ̇) of the reaction wheel, moment of inertia of the spacecraft (Ispacecraft),
the angle by which the spacecraft manoeuvres (ɵ) and the density of the wheel (ρ).
From Buckingham π theorem we can say that there are three dimensional groups.
They are

𝜋1 = ɵ

2
𝐼𝑠𝑝𝑎𝑐𝑒𝑐𝑟𝑎𝑓𝑡 𝑉𝑖𝑛 𝐾𝑉 2
𝜋2 =
ℎ̇

𝜋3 = 𝑡𝑉𝑖𝑛 𝐾𝑉

Hence we can say that


2
𝐼𝑠𝑝𝑎𝑐𝑒𝑐𝑟𝑎𝑓𝑡 𝑉𝑖𝑛 𝐾𝑉 2
𝑡𝑉𝑖𝑛 𝐾𝑉 = 𝑓(ɵ, ̇ℎ
) … (26)

We will try and vary all the variables and plot them in order to find the function that
relates them all.

14.0 References
Anon., n.d. CubaSatShop. [Online]
Available at: https://www.cubesatshop.com/product/crystalspace-p1u-vasik/
[Accessed 17 11 2016].

Anon., n.d. Cubesat. [Online]


Available at: http://www.cubesat.org/
[Accessed 9 12 2016].

Anon., n.d. diydrones. [Online]


Available at: http://diydrones.com/profiles/blogs/motor-efficiency
[Accessed 22 11 2016].

Anon., n.d. Faulhaber. [Online]


Available at:
https://fmcc.faulhaber.com/details/overview/PGR_4563_13825/PGR_13825_13814/en/SG/
[Accessed 4 12 2016].

27
Anon., n.d. Learningc. [Online]
Available at: http://learningrc.com/motor-kv/
[Accessed 29 11 2016].

Anon., n.d. Taxxas. [Online]


Available at: https://traxxas.com/forums/showthread.php?202168-Efficiency-curve-of-
brushless-vs-brushed
[Accessed 6 10 2016].

Auret, J., 2012. Design of an Aerodynamic Attitude Control System for a CubeSat, Matieland:
s.n.

Cannon, R. H., n.d. Some Basic Response Relations For Reaction Wheel Attitude Control,
Stanford: s.n.

Garza, A. M. C., 2008. Reaction Wheels for Picosatellites, Kiruna: s.n.

Herl, M., 2010. Estimation of Magnetic Torquing and Momentum Storage Requirements for
a Small Satellite, Colorado: s.n.

Liu, L., 2007. [Online]


Available at:
https://www.google.co.uk/url?sa=t&rct=j&q=&esrc=s&source=web&cd=1&cad=rja&uact=8
&ved=0ahUKEwj484e6tPfQAhVEJ8AKHRNcDkQQFggaMAA&url=http%3A%2F%2Fweb.mit.ed
u%2Flululiu%2FPublic%2FTESS%2520things%2Facs_analysis.pdf&usg=AFQjCNFHNg-
JOD7u8HnDzGTDk9osGZIGVg&bvm=b
[Accessed 22 10 2016].

Long, F., 2014. Design and testing of a Naosatellite Simulator Reaction Wheel Attitude
Control System, Utah: s.n.

Markley, F. L. & Crassidis, J. L., 2014. Fundamentals of Spacecraft Attitude Determination


and Control, New York: Springer.

Masterson, R. A., 1997. Development and validation of Empirical and Analytical Reaction
Wheel Disyurbance Models, s.l.: s.n.

Michaelis, T., n.d. Small Satellite Reaction Wheel Optimisation, Maryland: s.n.

28
O'Hearn, T. L., n.d. Calculation Motor Start Time, s.l.: s.n.

Oland, E. & Sclanbusch, R., 2009. Reaction Wheel Design for Cubesats, Narvik: s.n.

Ousaloo, H. S., 2016. Active nutation control of an asymmetric spacecraft using an reaction
wheel. ScienceDirect, 27 7.

Snider, R. E., 2010. Attitude Control of a Satellite Using Reaction Wheels and a PID
Controller, s.l.: s.n.

Wertz, J. R., Everett, D. F. & Puschell, J. J., 2011. Space Mission Engineering: The New SMAD.
1 ed. Hawthorne: Microcosm press.

Zeledon, R. A. & Peck, M. A., n.d. Attitude Dynamics and Control of a 3U CubeSat with
Electrolysis Propulsion, New York: s.n.

15.0 Appendix

xx xy xz xy yy yz xz zy zz
-2.9E- 1.756E- 8.78E-
1 0.001528 1.756E-05 06 05
0.001454 06 -2.9E-06 8.78E-06 0.001545
-5.9E- -8.78E- -8.8E-
2 0.001481 -8.78E-06 06 060.00146 06 -5.9E-06 -8.8E-06 0.001492
2.93E- -2.6E-
3 0.001598 0 06 0 0.001454 05 2.93E-06 -2.6E-05 0.001615
2.93E- -8.78E- -3.5E-
4 0.001893 -8.78E-06 06 06 0.001454 05 2.93E-06 -3.5E-05 0.001911
5.85E- -2.05E- -2.9E-
5 0.001612 -2.05E-05 06 05 0.00146 05 5.85E-06 -2.9E-05 0.001624
-5.85E- 2.63E-
6 0.001329 -5.85E-06 0 06 0.001457 05 0 2.63E-05 0.001343
2.93E- 5.853E- 4.39E-
7 0.001451 5.853E-06 06 06 0.001457 05 2.93E-06 4.39E-05 0.001466
1.463E- -1.8E-
8 0.001744 1.463E-05 0 05 0.001454 05 0 -1.8E-05 0.001762
-3.51E- 1.46E-
9 0.001551 -3.51E-05 0 05 0.001454 05 0 1.46E-05 0.001569
2.93E- 5.853E- 5.85E-
10 0.001323 5.853E-06 06 06 0.001454 06 2.93E-06 5.85E-06 0.00134
5.85E- 1.756E- 4.39E-
MAX 0.001893 1.756E-05 06 05 0.00146 05 5.85E-06 4.39E-05 0.001911
-5.9E- -3.51E- -3.5E-
MIN 0.001323 -3.51E-05 06 05 0.001454 05 -5.9E-06 -3.5E-05 0.00134
8.78E- -3.51E- -1.8E-
Average 0.001551 -3.51E-06 07 06 0.001456 06 8.78E-07 -1.8E-06 0.001567
Appendix 1. Inertia matrix tabulation for all the possible mass distributions (all in SI units).

29

Das könnte Ihnen auch gefallen