n mcnamara - espace.library.uq.edu.au

63

Upload: others

Post on 11-Dec-2021

27 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: N McNamara - espace.library.uq.edu.au
Page 2: N McNamara - espace.library.uq.edu.au

N McNamara

i

Abstract

This report investigates and compares steering laws for a control moment gyroscope (CMG)

attitude control system for a spacesuit. If there is an inadequate level of control in a CMG

system then the angular momentum vectors of individual CMGs may become aligned. This

state, referred to as a β€˜singularity state’, prevents the system from being able to produce

torque in one or more directions. The steering law problem has been defined and the

momentum envelope is presented to visualize system singularities. A literature review has

been conducted to find applicable steering laws that may be used to control a spacesuit sized

system and avoid singularities in the momentum envelope. Six potential steering laws are

identified as being potentially applicable to a spacesuit sized system; the Moore-Penrose

pseudoinverse (MPI), offline planning, preferred gimbal angle, linearly constrained, null

motion, and the singularity-robust inverse (SRI).

A conceptual pyramid configuration system of 4 CMGs has been constructed to analyse the

selected steering laws through a qualitative and quantitative comparison. The qualitative

assessment investigates several criteria; the singularity handling method, induced torque

error, system efficiency and full utilization of hardware, convergence to singularity

arrangements, and the potential for generalization to other CMG arrangements. The results

of the qualitative comparison conclude that the SRI control law is the most viable for use with

a spacesuit. A computational model was used to compare the behaviour of the conceptual

system using the MPI and the SRI control laws. The simulation results support the qualitative

analysis. The SRI control law successfully navigates around singularities by inducing a small,

proportional torqueing error. It is concluded that a CMG control system can potentially be

used with a spacesuit and that the SRI control law is the most viable method to avoid

singularities during operation.

Page 3: N McNamara - espace.library.uq.edu.au

N McNamara

ii

Page 4: N McNamara - espace.library.uq.edu.au

N McNamara

iii

Contents Abstract ..................................................................................................................................................... i

List of Tables and Figures ......................................................................................................................... v

1. Introduction .................................................................................................................................... 1

1.1. Motivation and Problem Definition ........................................................................................ 3

1.2. Astronaut Musculoskeletal Injury Rates ................................................................................. 4

1.3. Project Aims ............................................................................................................................ 5

2. Scope ............................................................................................................................................... 6

3. Literature Review ............................................................................................................................ 7

3.1. Overview of Microgravity Attitude Control and Dynamics ..................................................... 7

3.2. Propellant Based Systems ....................................................................................................... 7

3.3. Gyroscopic Attitude Control Systems ..................................................................................... 8

3.3.1. Torque Reaction Wheels ..................................................................................................... 9

3.3.2. Control Moment Gyroscopes .............................................................................................. 9

3.3.3. Vehicle Characteristics and Gyroscopic Control System Efficiency ................................... 11

3.4. CMG System Arrangements and Singularities ...................................................................... 12

3.4.1. Singularity Avoidance and Null Motion ............................................................................. 15

4. Project Methodology and Process ................................................................................................ 16

5. System Model ................................................................................................................................ 17

5.1. Dynamics and Equations of Motion ...................................................................................... 17

5.2. Pseudoinverse and Minimum Norm Solution ....................................................................... 22

5.3. Momentum Envelope and CMG Gain ................................................................................... 24

5.4. Elliptical and Parabolic Singularities ...................................................................................... 25

Page 5: N McNamara - espace.library.uq.edu.au

N McNamara

iv

6. Steering Laws ................................................................................................................................. 29

6.1. No Singularity Avoidance ....................................................................................................... 31

6.1.1. Moore-Penrose Pseudoinverse ......................................................................................... 31

6.2. Singularity-Avoidance ............................................................................................................ 32

6.2.1. Offline Planning ................................................................................................................. 32

6.2.2. Preferred Gimbal Angle/Gimbal Reorientation ................................................................. 33

6.2.3. Linearly Constrained .......................................................................................................... 34

6.3. Singularity-Robust ................................................................................................................. 36

6.3.1. Gradient/Null Motion Weighted Matrix ............................................................................ 36

6.3.2. Singularity-Robust Inverse ................................................................................................. 37

6.4. Qualitative Summary ............................................................................................................. 39

7. Simulation and Results .................................................................................................................. 40

8. Discussion ...................................................................................................................................... 43

9. Conclusion ..................................................................................................................................... 45

10. References ................................................................................................................................. 46

11. Appendices ................................................................................................................................ 49

11.1. Appendix A – CMG Torque Derivation from First Principles ............................................. 49

11.2. Appendix B – Simulation Data ........................................................................................... 51

Page 6: N McNamara - espace.library.uq.edu.au

N McNamara

v

List of Tables and Figures

Figure 1 - Astronaut affixed to Canadarm2 on the ISS ............................................................................ 2

Figure 2 - Astronaut with MMU attached ............................................................................................... 4

Figure 3 - Attitude errors, propellant-only attitude control system (left), N number of CMG controllers

added to system (right) (Source: Carpenter, et al., 2013) ...................................................................... 8

Figure 4 - Reaction wheel assembly momentum vectors ..................................................................... 10

Figure 5 - Single-gimbal CMG momentum vectors ............................................................................... 10

Figure 6 - RWA and CMG comparison of momentum potential envelopes (Source: Hamilton, et al.,

2015) ..................................................................................................................................................... 11

Figure 7 - CMG pyramid configuration .................................................................................................. 12

Figure 8 - CMG roof configuration ........................................................................................................ 13

Figure 9 - Control moment gyroscope diagram (Source: Yoon, 2004) .................................................. 17

Figure 10 - Pyramid CMG system diagram ............................................................................................ 18

Figure 11 - Momentum envelope of conceptual system (Source: Yoon, 2004) ................................... 25

Figure 12 - Hyperbolic singularity momentum envelope surface (Source: Yoon, 2004) ...................... 26

Figure 13 - Hyperbolic singularity CMG arrangement .......................................................................... 26

Figure 14 – Parabolic singularity momentum envelope surface (Source: Yoon, 2004) ........................ 27

Figure 15 - Parabolic singularity CMG arrangement ............................................................................. 27

Figure 16 - Momentum envelope of pyramid configuration CMG system (Source: Yoon, 2004) ........ 28

Figure 17 – Constrained steering law applied to system of 3 CMGs (Source: Jones, et al., 2012) ....... 35

Figure 18 - Effects of constrained steering law on momentum envelope (Source: Kurokawa, 1997) . 35

Figure 19 - Comparison of MPI and SRI Solutions ................................................................................. 42

Figure 20 - SRI Solution Induced Torque Error from Input Command .................................................. 42

Table 1 - Scope of the project ................................................................................................................. 6

Table 2 - Qualitative comparison summary .......................................................................................... 39

Page 7: N McNamara - espace.library.uq.edu.au
Page 8: N McNamara - espace.library.uq.edu.au

N McNamara

1

1. Introduction

Control moment gyroscopes (CMGs) are momentum-exchange actuators which are used to

control the attitude of spacecraft’s and other terrestrial vehicles. Gyroscopic control systems

are especially useful for space environments due to the independence of the system from

fuel-based energy sources and propellants. Using electrics motors and actuators a gyroscopic

system can be used indefinitely provided that adequate electrical power be provided, the

most prevalent source is solar energy.

Arrays of CMGs can be used in tandem to control a system in multiple planes of motion.

Different configurations of these arrays present various benefits and restrictions depending

on the applicable scenario. There are various methods of controlling these CMGs. The control

system must account for both the degree of actuation possible for a single CMG, and the

relative moment of inertia between CMGs in an array. If there is an inadequate level of control

in a CMG system then the angular momentum vectors of individual CMGs may become

aligned. This state, referred to as a β€˜singularity state’, prevents the system from being able to

produce torque in one or more directions. An analog can be made between the singularity

state of a CMG system and that of a mechanical system, for example a delta-type parallel

manipulator (Bedrossian, Paradiso, & Bergmann, 1990).

Mechanical Manipulator

CMG System

Position 𝒔 = 𝒔(𝒒) Momentum 𝒉 = 𝒉(𝜽) Velocity οΏ½Μ‡οΏ½ = 𝐽(𝒒)οΏ½Μ‡οΏ½ Torque 𝝉 = 𝐽(𝜽)οΏ½Μ‡οΏ½ Acceleration �̈� = 𝐽(𝒒)�̈� + 𝐽(𝒒)οΏ½Μ‡οΏ½ Rotatum (

𝑑

𝑑𝑑 Torque) οΏ½Μ‡οΏ½ = 𝐽(𝜽)�̈� + 𝐽(𝜽)οΏ½Μ‡οΏ½

System singular when no motion possible in a certain direction.

System singular when no torque possible in a certain direction.

Page 9: N McNamara - espace.library.uq.edu.au

N McNamara

2

Modern day astronaut suits employ propellant based systems for principle motion and

attitude control during EVA missions. With the world moving rapidly towards spaceflight and

operations further away from Earth, spacesuits need to be lighter, more mobile, and safer

than what is currently available. The design and research into spacesuit technology has be

largely stagnant since the shuttle program. Current extravehicular activity (EVA) missions are

only carried out near a vehicle in microgravity and thus astronauts primarily rely on operator

controlled robot arms connected to space vehicles (Figure 1), and manual movement about a

vehicle by grabbing onto external handles to conduct operations in outer space.

When astronauts are required to venture further from the parent space vehicle, external

booster packs are affixed to an EVA suit. Propellant is used for principle motion and for

attitude control of the astronaut. The use of propellant for attitude control reduces the

available mission time of the suit for a given mass, and increases the mass of the system

unnecessarily. A new form of attitude stabilization is required in such a scenario.

Figure 1 - Astronaut affixed to Canadarm2 on the ISS

Page 10: N McNamara - espace.library.uq.edu.au

N McNamara

3

1.1. Motivation and Problem Definition

The development of the spacesuit has been predominantly stagnant since the inception of the

Apollo program in 1967 (Harland, 2010). Spacesuits are categorized into operational areas and

provide different functionality depending on the scenario. The Extravehicular Mobility Unit

(EMU) was developed as an independent anthropomorphic spacesuit for deployment in outer-

and low-atmospheric environments (Thomas & McMann, 2006). The EMU system provides

environmental protection, mobility, life support, and communications to operators, an EMU

is essentially the smallest human space vehicle ever created (West & Witt, 2010). During

extravehicular activities astronauts use the EMU in conjunction with an array of separate,

modular units which provide different modes of position and attitude control in zero gravity.

The main unit used is the Manned Manoeuvring Unit (MMU), shown in Figure 2. The MMU

attaches to the back of the EMU and is worn like a backpack. The unit consists of a small one-

man nitrogen-propellant based propulsion system that the user can use for attitude control

and translational motion (Cheng, 2010). Counterintuitively, during a spacewalk an astronaut

does very little with their legs. The principle mode of traversing the exterior of a space vehicle

is by way of manually clambering along the hull using hand holds and bracing. In the vacuum

of space the pressurized, rubber composite spacesuit gloves expand and harden making

simple tasks difficult to perform. Current design constraints and a lack of alternative spacesuit

designs force astronauts to perform EVA operations in far more difficult circumstances than

immediately apparent (Danaher, Tanaka, & Hargens, 2005). When fine motor control is

required to complete tasks during an operation an astronaut must constantly control their

attitude and this inevitably leads to errors when dealing with precision tooling or scientific

sample collection.

Page 11: N McNamara - espace.library.uq.edu.au

N McNamara

4

Figure 2 - Astronaut with MMU attached

1.2. Astronaut Musculoskeletal Injury Rates

One of the most prominent fields of study relating to human spaceflight is the potential risk

of musculoskeletal injury and minor trauma, more specifically, in the hands and forearms, of

astronauts (Opperman, et al., 2010). Jennings and Bagian’s (1996) study investigating

terrestrial-based musculoskeletal injuries of astronauts in the period of 1987 – 1995 found

astronauts sustained numerous minor injuries because of EVA-specific operations and

training. A total of 28 orthopedic surgical procedures were conducted in this period due to

sustained fractures and serious ligament, cartilage, and soft tissue injuries (Jennings & Bagian,

1996). Furthermore, post-flight medical debrief statistics between the period of 1996 – 2006

of all American and Russian spaceflight operations found a total of 219 in-flight

musculoskeletal injuries (Scheuring, et al., 2009). The study found that hand and forearm

Page 12: N McNamara - espace.library.uq.edu.au

N McNamara

5

injuries dominate throughout the space program accounting for 47% of spacewalk related

injuries in the period of 2002 – 2004. It was also shown that EVA workers have a 19.6% chance

of incurring an injury during operations. The most common injuries include fingernail

delamination, minor muscle tears, tendon shortening, and minute fractures (Opperman,

Waldie, Natapoff, Newman, & Jones, 2010).

Current injury rates of astronauts are directly correlated with EVA operational practices. The

manual traversing and stabilization of astronauts about the parent space vehicle results in

numerous medical issues and hinders further operations (Danaher, Tanaka, & Hargens, 2005).

A new form of stabilization which will allow astronauts to work unhindered in zero-gravity

vacuum environments while maintaining a given attitude is required.

1.3. Project Aims

Before a new method of attitude control can be implemented into operation on astronaut

EVA suits, the behavioural characteristics of various passive and reactionary systems that

could potentially fill the role must be understand. Current literature on the topic of gyroscopic

attitude control falls outside the size-envelope of a one-man, spacesuit-sized system. As will

discussed below in Section 3, micro- and larger scale satellites use gyroscopes for attitude

control. These systems, however, are not applicable to a manned unit. More specifically the

various control laws used to control the system of CMGs on these vehicles must be

investigated for viability for use with a spacesuit. The focus of this thesis is to contribute to

the advancement of knowledge in the field of astronautics and system dynamics engineering.

The aim of the project is the qualitative and quantitative comparison of control moment

gyroscope system control laws for use with a spacesuit sized vehicle.

Page 13: N McNamara - espace.library.uq.edu.au

N McNamara

6

To evaluate the performance and applicability of gyroscopes for the stated purpose the

project will investigate several topics to develop a final proposed system. The investigated

topics include;

- Zero gravity astronautics and dynamics of EVA suit sized systems,

- Propellant based attitude control packages,

- The functionality and characteristics of control moment gyroscopes, and

- The momentum envelope capabilities of CMG configurations.

2. Scope

Table 1 details aspects of the project which are inside and outside of the scope for this thesis.

Table 1 - Scope of the project

In Scope;

- Investigation of current technology and components.

- Computational models of the system using software packages such as Matlab, Simulink,

Solidworks, Ansys, and Python.

- Quantitative and qualitative comparison of CMG system control laws as found in literature.

- Theoretical and computer modelling of the proposed system.

Out of Scope;

- Bespoke software package coding or extensive computational work that requires a power

unit greater than that of a consumer grade PC.

- Computational fluid dynamics on rotor effects, these are assumed to be negligible given that

the rotors are enclosed in a small housing in zero gravity.

- Experimentation on designed system/s.

- Prototyping of designed components.

- Development of new steering law theory.

Page 14: N McNamara - espace.library.uq.edu.au

N McNamara

7

3. Literature Review

3.1. Overview of Microgravity Attitude Control and Dynamics

Modern satellites and space vehicles use a range of methods for orientation and momentum-

management. There are three industry accepted technologies which allow this; high-torque

Reaction Wheel Assemblies (RWA), Control Moment Gyroscopes (CMG), and propellant based

systems (Leve, Hamilton, & Peck, 2015). The first two technologies are forms of energy and

momentum management within a system, the latter is a form of reactionary-control and

results in a non-zero net force on the system. Gyroscopic systems are continuous and present

a much finer degree of control as compared to the discrete propellant-based systems.

Although propellant based-systems can be used for small attitude adjustments the discrete

nature of the impulses from the jets result in a constant oscillation about the desired attitude

vector. When compared to gyroscopic systems there is a constant attitude error when using

propellant systems, as illustrated in Figure 3. The system of equations used to describe each

type of control is therefore different (Tewari, 2011). Unlike atmospheric flight vehicles,

attitude and orbital dynamics of a space vehicle are uncoupled allowing for independent

design and analysis (Leve, Hamilton, & Peck, 2015). This report deals exclusively with the

design and investigation into attitude control for a spacesuit. Translation and attitude-

translation combination manoeuvres fall outside the scope.

3.2. Propellant Based Systems

Propellant based systems have been used since the inception of the space program. Consisting

of a number of gimbaled or fixed nozzles, propellant is ejected from the parent body in order

to change or control the attitude of the system. This form of control is referred to as a reaction-

Page 15: N McNamara - espace.library.uq.edu.au

N McNamara

8

control system whereby the system must react with another body or its surrounds to create a

change in the momentum of the spacecraft. Carpenter, et al. (2013) modelled an experimental

MMU with additional attitude control specific jet units. The designed system was then

analyzed with and without the addition of a set of 𝑛 control moment gyroscopes. The results

show that the propellant-control only system developed large perturbations and uncertainties

when stabilizing the system to a set attitude. The propellant jets result in faster slew rates

however the total manoeuvre time without CMGs is extended due to the convergence rate

(Carpenter, et al., 2013).

Figure 3 - Attitude errors, propellant-only attitude control system (left), N number of CMG controllers added to system (right) (Source: Carpenter, et al., 2013)

3.3. Gyroscopic Attitude Control Systems

Reactionary systems require the use of finite propellant and therefore present large

inefficiencies for missions over ever-increasing duration and distance from Earth. Managing a

spacecraft’s momentum with rotors allows for the system to be controlled for an infinite

period given the required energy to maintain sufficient velocity of the rotors. There are two

different ways of managing the momentum of a system using gyroscopes, torque reaction

wheel assemblies (RWAs) and control moment gyroscopes (CMGs). The first uses a change in

flywheel velocity to enact a desired torque on the system, whereas the latter uses a changes

Page 16: N McNamara - espace.library.uq.edu.au

N McNamara

9

the spin-moment vector of a constant speed fly-wheel to achieve the same. Both technologies

are discussed further in the next sections. A combination of the two technologies can also be

used, a variable speed control moment gyroscope (VSCMG). VSCMG systems are mentioned

in this report however they fall outside the scope of the project.

3.3.1. Torque Reaction Wheels

An RWA consists of a brushless motor attached to a high-inertia flywheel. The vector direction

of the flywheel is fixed about the spacecraft. The magnitude of the momentum vector varies

as the flywheel is sped up or slowed down. The mechanical shaft power is equal to the torque

times the speed of the rotor: 𝑷 = 𝝉 βˆ™ 𝝎 (Leve, et al., 2015). Therefore, the resultant torque

applied to a spacecraft (𝝉𝑺π‘ͺ) because of input power 𝑷 is thus; 𝝉 =𝑷

𝝎= βˆ’π‰π‘Ίπ‘ͺ. This type of

control can be viewed as a form of β€˜momentum exchange’ whereby the momentum of the

flywheel is either increased or decreased by using electrical motors. The momentum energy

is transferred to electrical energy and stored, or vice versa. The spacecraft then experiences

an equal and opposite torque to that imparted on/by the flywheel. Although ideally this type

of system would be entirely reversible inefficiencies are introduced when converting work to

energy, and energy to work. Large energy storage capacity and electrical motors are also

required in addition to the flywheel.

3.3.2. Control Moment Gyroscopes

A CMG utilizes a secondary axis, or tertiary axis in some cases, to change the direction of the

momentum vector. In most applications, a CMG consists of a relatively constant speed

flywheel, however, variable-speed systems have become more prevalent (Leve, et al., 2015).

Page 17: N McNamara - espace.library.uq.edu.au

N McNamara

10

The spacecraft and the CMG system exchange momentum to control and/or maintain the

attitude of the spacecraft. In contrast to an RWA system, the momentum of the entire system

is always maintained. Small amounts of energy are used to drive the gimbal motors however

the momentum of the flywheel and spacecraft as a whole is constant. The rotor with angular

momentum vector 𝒉, with magnitude 𝑱𝒓𝛀𝒓 precesses dependent on applied torque 𝝉, such

that 𝝉 = 𝝎 Γ— 𝒉 = βˆ’π‰π‘Ίπ‘ͺ (Leve, et al., 2015).

Figure 4 - Reaction wheel assembly momentum vectors

Figure 5 - Single-gimbal CMG momentum vectors

𝜏

𝜏

πœ”

Page 18: N McNamara - espace.library.uq.edu.au

N McNamara

11

3.3.3. Vehicle Characteristics and Gyroscopic Control System Efficiency

As stated in Section 3.3.1 the torque potential of an RWA is limited by the mechanical power

of the motor. CMGs do not face the same restrictions. Hamilton, et al. (2015) presents the

differing capability envelopes of RWA, single-gimbal CMG, and double-gimbal CMG designs

(Figure 6). Similarly, Votel and Sinclair (2013) conducted evaluations of commercially available

gyroscopic control systems for small scale satellites. The authors present the results of their

analysis comparing the power consumption contours for each system type dependent on the

slew angle potential/requirements of the spacecraft. They concluded that for any satellite

exceeding a mass of 30kg should strongly consider a CMG system, and above 100kg there is

very little choice but to use a CMG (Votel & Sinclair, 2013).

Figure 6 - RWA and CMG comparison of momentum potential envelopes (Source: Hamilton, et al.,

2015)

Page 19: N McNamara - espace.library.uq.edu.au

N McNamara

12

3.4. CMG System Arrangements and Singularities

There are infinite number of CMG configurations possible for a given system, these can be

categorized based on the net angular momentum of the system in the nominal position (Leve,

Hamilton, & Peck, 2015). More modern CMG configurations include the pyramid (Figure 7)

and scissor (or roof) (Figure 8) configurations. The configuration of the CMGs changes the

momentum envelope potential of the system, that is, the maximum potential torque that can

be applied in a certain direction. The details of the momentum envelope for a given system

are investigated mathematically below in Section 5.3. Although these configurations are more

generally used with single-gimbal CMGs, the limiting saturation alignment profiles are also

used in dual-gimbal CMG systems (Leve, Hamilton, & Peck, 2015).

Figure 7 - CMG pyramid configuration

Page 20: N McNamara - espace.library.uq.edu.au

N McNamara

13

Figure 8 - CMG roof configuration

A singularity refers to the instance where two-or-more CMGs in a system have aligned or

exactly opposite momentum vectors. For systems with a cluster of 𝑁 CMGs, when 𝐾 CMGs

are in a singular state then the system is only capable of producing torque in (𝑁 βˆ’ 𝐾)

dimensions. That is to say; for a common system of 4 CMGs, if 2 CMGs are in a singular state

then the system is only capable of producing a torque in 2 dimensions (McMahon & Schaub,

2009).

If the system is not controlled adequately then singularities within a CMG system can lead to

limited operational capabilities, total inoperability, and ultimately, system failure. Consider

the case where the required spacecraft attitude change results in an ever-increasing storage

of momentum in the CMG system. As momentum is exchanged between the spacecraft and

the gyroscopic control system the CMGs approach the maximum angular momentum state,

the point at which no further angular momentum can be stored in the system. Figure 6

illustrates this theoretical circumstance. Starting in a net-zero state where all CMGs have equal

Page 21: N McNamara - espace.library.uq.edu.au

N McNamara

14

and opposite momentum vectors, and progressing to a saturation state where all momentum

vectors are aligned.

The system must then be desaturated to allow the CMG control system to dissipate angular

momentum. A torque must be applied to the spacecraft which the CMG system can react

against to realign the gyroscopes into the nominal position. This can be achieved by firing

propellant thrusters, deploying a magnetic torque rod which will react against the Earth’s

magnetic field, or any other means which produces an outside torque on the system (Russel,

Spencer, & Metrocavage, 2008).

The saturation limit is dictated by component and configuration characteristics and exceeding

this limit can cause catastrophic failure (Kim, 2011). The International Space Station (ISS) uses

a system of four dual-gimbal CMGs with a life expectancy of 10 years and an output torque of

258 Nm each (Gurrisi, et al., 2010). Two of the gyroscopes, designated CMG1 and 3, failed

prematurely. Gurrisi, et al. (2010) conducted a failure analysis and determined that CMG1 had

a β€˜hard’, catastrophic failure due to excessive lateral forces on the gimbal bearing, CMG3 had

a β€˜soft’ failure due to fretting on a bearing sleeve that resulted in excessive perturbation and

error. The initial failure began because of a large angular-momentum transfer prior to a

Shuttle docking procedure. Although manufacturing and assembly defects were the cause of

the failure the excessive strain on the supporting structure due to the saturated-nominal

position aggravated the defects (Gurrisi, et al., 2010).

Page 22: N McNamara - espace.library.uq.edu.au

N McNamara

15

3.4.1. Singularity Avoidance and Null Motion

Avoiding singularities has been at the forefront of astronautic research in the recent past as

satellites become ever-smaller and the need for a control system utilizing three or less CMGs

becomes more apparent. Several authors have investigated steering laws for constant-speed

CMG clusters and other authors have investigated the potential of variable-speed CMG

(VSCMG) clusters. The avoidance of singularity states is achieved by means of either efficiency

path-planning, or null motion. Path-planning is widely used on satellites when attitude control

is pre-determined before-the-fact. This method of singularity avoidance is not applicable to a

spacesuit given the user operated nature of the conceptual system. Null motion refers to the

controlled momentum cancelling of two or more CMGs against each other.

Imagine that a user controlled manoeuvre results in two CMGs approaching a singularity state.

As is the case with a constant speed, single-gimbal CMG system, there is no way to increase

or decrease the momentum of any singular gyroscope. Only the gimbal angles can be

controlled by actuating the gimbal motors. The only way to avoid a singularity is to react the

individual gyroscopes against each other in a manner which results in a net-zero torque on the

system. In plain terms, two or more gimbals are actuated slightly to β€˜bleed’ some momentum

energy away from the system, the result of which is the avoidance of a singularity while the

overall system can achieve the desired attitude change. The term null motion refers to the

comparative motion between the individual CMGs in a manner that results in a net-zero, null,

torque on the system.

Page 23: N McNamara - espace.library.uq.edu.au

N McNamara

16

4. Project Methodology and Process

As outlined in Section 1.3, the aim of the project is the qualitative and quantitative comparison

of CMG system control laws for use with a spacesuit sized vehicle. To achieve this the project

has been broken into two separate stages, an initial qualitative comparison through literature,

and a quantitative comparison using a Python model of a conceptualized system. The

following steps outline the methodology used to achieve the desired project outcomes and to

meet the stated aim;

1. Defined the generalized CMG system from first principles.

2. Designed a conceptual system that will be used to compare steering laws.

3. Defined the singularity problem and thus the steering law problem, mathematically.

4. Found steering laws in the literature that address the steering law problem and that

are applicable to the conceptual system.

5. Investigated the chosen steering laws individually for qualitative characteristics and

compared.

6. Using the defined conceptual system, designed a Python model that computes the

system response for different control laws.

7. Implemented the different control laws being tested into the model.

8. Tested the modelled system response for a user-defined attitude manoeuvre and

compared the results.

9. Made an overall conclusion on the most appropriate control law for a spacesuit CMG

system and made comments regarding potential changes that would be required for

implementation.

Page 24: N McNamara - espace.library.uq.edu.au

N McNamara

17

5. System Model

5.1. Dynamics and Equations of Motion

Consider a system with a single CMG device as depicted in Figure 9. By torqueing the gimbal,

the flywheel orientation is changed, thereby re-directing the rotor’s angular momentum. The

derivation of Equation 1 from first principles is shown in Appendix A. The resultant torqueing

on the overall system is equal and opposite to that imparted by the internally mounted CMG;

𝝉𝑖 = 𝒉𝑖 Γ— 𝛾�̇� = βˆ’π‰π‘†πΆ

where;

𝝉𝑖 = Torqueing vector of CMG 𝑖

π‰π‘†π‘Œπ‘† = Spacesuit torqueing vector

π’‰π‘Ÿπ‘–= Flywheel angular momentum vector

𝛾�̇̇� = CMG gimbal angle actuation rate

Figure 9 - Control moment gyroscope diagram (Source: Yoon, 2004)

(1)

Page 25: N McNamara - espace.library.uq.edu.au

N McNamara

18

Jones, Zeledon & Peck (2012) investigated potential configurations for single-gimbal CMG

systems. The authors concluded that for a system with constant speed-flywheels, 𝑁 CMGs is

the most efficient system size for (𝑁 βˆ’ 1)-dimensional attitude control. Therefore, a

conceptual system consisting of 4 CMGs has been used for this report. Equation 1 can be

adapted for a system of 𝑁 = 4 CMGs;

𝝉TOT = βˆ‘(𝒉𝑖 Γ— 𝛾�̇�)

𝑁=4

𝑖=1

Applying the condition that all CMGs be identical in moment of inertia and angular velocity

the system model allows for simplification in design and modelling. Consider the conceptual

system depicted in Figure 10. The system has been designed using a typical pyramid

configuration as discussed in Section 3.4.

Figure 10 - Pyramid CMG system diagram

(2)

𝒉1

𝒉2

𝒉3

𝒉4 𝜸1

𝜸2

𝜸3

𝜸4

𝑀

𝛽

Page 26: N McNamara - espace.library.uq.edu.au

N McNamara

19

The system model must account for several angular momentum components related to;

β€’ Point mass of the spacesuit,

β€’ Moment of inertia of the spacesuit,

β€’ Angular velocity of the spacesuit,

β€’ Point masses of the CMGs,

β€’ Flywheel angular velocity, and

β€’ Gimbal angular velocity.

The rotational equation of motion of a rigid spacecraft equipped with a momentum-exchange

device can be described as;

οΏ½Μ‡οΏ½ + 𝝎 Γ— 𝑯 = 𝑻EXT

where 𝑯 = (𝐻π‘₯, 𝐻𝑦 , 𝐻𝑧) is the angular momentum vector of the spacesuit relative to the

origin, 𝝎 = (πœ”1, πœ”2, πœ”3) is the spacesuit angular velocity vector, and 𝑻EXT is the external

torque vector applied to the spacesuit system (Wie, 2001). The cross product 𝝎 Γ— 𝑯 in matrix

notation takes the form;

πŽΓ—π‘― = [

0 βˆ’πœ”π‘§ πœ”π‘¦

πœ”π‘§ 0 βˆ’πœ”π‘₯

βˆ’πœ”π‘¦ πœ”π‘₯ 0] [

𝐻π‘₯

𝐻𝑦

𝐻𝑧

]

Assuming that the system is rigid and that there are negligible losses in energy between the

CMGs and the parent spacesuit, the equation for angular momentum by Gui, Vukovich & Xu

(2016) can be used. The authors state that the angular momentum of the system about centre

of mass, M, for a system with 𝑁 CMGs can be described as the sum of the spacecraft angular

moment and that of the individual CMGs;

𝑯 = π‘±πŽ + 𝐴𝑔𝐼𝑐𝑔�̇� + 𝐴𝑠𝐼𝑀𝑠𝛀

(3)

(4)

Page 27: N McNamara - espace.library.uq.edu.au

N McNamara

20

Where 𝜸 = (𝛾1, … , 𝛾𝑁)𝑇 ∈ ℝ𝑁 and 𝛀 = (Ξ©1, … , Ω𝑁)𝑇 ∈ ℝ𝑁 represent the gimbal angles and

flywheel speed vectors of each CMG in the individual gimbal reference frames, respectively.

The matrix 𝐽 is the Jacobian inertia matrix for the whole spacesuit system defined as;

𝑱 = 𝐼𝐡 + 𝐴𝑠𝐼𝑐𝑠𝐴𝑠𝑇 + 𝐴𝑑𝐼𝑐𝑑𝐴𝑑

𝑇 + 𝐴𝑔𝐼𝑐𝑔𝐴𝑔𝑇

Where 𝐼𝐡 is the inertia matrix of the spacesuit system and the point masses of each CMG. The

matrices πΌπ‘βˆ— and πΌπ‘€βˆ— are diagonal with elements of inertia for the gimbal plus flywheel

structure and flywheel-only of each CMG. Therefore, πΌπ‘βˆ— = πΌπ‘€βˆ— + πΌπ‘”βˆ— where πΌπ‘€βˆ— =

diag[πΌπ‘€βˆ—1, … , πΌπ‘€βˆ—π‘] and πΌπ‘βˆ— = diag[πΌπ‘βˆ—1, … , πΌπ‘βˆ—π‘] for βˆ—= οΏ½Μ‚οΏ½, οΏ½Μ‚οΏ½ or οΏ½Μ‚οΏ½.

Matrices π‘¨βˆ— ∈ ℝ3×𝑁 represent the directional unit vectors in the roll (π‘₯), pitch (𝑦) and yaw (𝑧)

directions relative to the body-frame (as defined in Figure 10). Therefore, 𝑨𝑅 = [οΏ½Μ‚οΏ½1, … , �̂�𝑁],

𝑨𝑃 = [οΏ½Μ‚οΏ½1, … , �̂�𝑁], and π‘¨π‘Œ = [οΏ½Μ‚οΏ½1, … , �̂�𝑁], such that;

π‘¨βˆ— = [𝑨𝑅

𝑨𝑃

π‘¨π‘Œ

] = [

("Roll Axis Component")

("Pith Axis Component")

("Yaw Axis Component")]

Because the matrices π΄βˆ— components are dependent on the individual gimbal angles, 𝛾𝑖, ie

𝐴𝑔 = 𝐴𝑔(𝛾) and 𝐴𝑠 = 𝐴𝑠(𝛾), by definition π΄βˆ— = π΄βˆ—(𝛾) and therefore 𝑱 = 𝑱(𝛾). Note that the

inertial matrix 𝐼𝐡 is constant, ie 𝐼𝐡 β‰  𝐼𝐡(𝛾).

Equation 4 can be further simplified by assuming the angular momentum contribution from

the gimbal structure is negligible compared to that of the flywheel and the overall system such

that π΄βˆ— = 𝐴. Due to the comparative nature of the final system model this assumption does

not adversely affect the project results. Therefore, eliminating the angular momentum

contribution due to gimballing, (𝐴𝑔𝐼𝑐𝑔�̇�), Equation 4 becomes;

𝑯 = π‘±πŽ + βˆ‘ 𝒉𝑖

𝑁=4

𝑖

= π‘±πŽ + 𝒉1 + 𝒉2 + 𝒉3 + 𝒉4 = π‘±πŽ + 𝒉

(5)

(6)

Page 28: N McNamara - espace.library.uq.edu.au

N McNamara

21

Carpenter et al. (2013) outlines the time derivative of Equation 6 with the assumption that the

point mass change of the CMGs relative to the overall system is negligible. Although this

assumption may be appropriate for a larger scale spacecraft system, such as the International

Space Station (ISS), this assumption does not hold true for the conceptual system. With

implementation on a spacesuit sized system the point masses of the CMGs have an

attributable effect on the angular momentum of the system. The methodology of the authors

has been used to find the time derivative of the angular momentum function with

modification to account for the CMG point mass angular velocity.

Introducing 𝐡 and 𝑂 denoting the body and inertial origin frames of reference respectively

and taking the time derivative;

𝑯 = π‘±Γ—πŽπ΅/𝑂 + βˆ‘ 𝒉𝑖

𝑁=4

𝑖

οΏ½Μ‡οΏ½ = 𝑱�̇�𝐡/𝑂 + βˆ‘ 𝒉𝑖̇

𝑁=4

𝑖

where 𝝎𝐡/𝑂 denotes the system angular velocity vector, 𝝎, in the body frame relative to the

inertial origin frame. The angular momentum of the CMGs for the conceptual system is;

𝒉 = βˆ‘ 𝒉𝑖(𝛾𝑖)

𝑁=4

𝑖=1

= [βˆ’cos 𝛽 sin 𝛾𝑖

cos 𝛾1

sin 𝛽 sin 𝛾1

] β„Ž1 + [

βˆ’ cos 𝛾2

βˆ’ cos 𝛽 sin 𝛾2

sin 𝛽 sin 𝛾2

] β„Ž2 + [cos 𝛽 sin 𝛾3

βˆ’ cos 𝛾3

sin 𝛽 sin 𝛾3

] β„Ž3 + [

cos 𝛾4

cos 𝛽 sin 𝛾4

sin 𝛽 sin 𝛾4

] β„Ž4

Noting that the CMG flywheels have equal angular velocity, β„Ž = β„Ž1 = β„Ž2 = β„Ž3 = β„Ž4;

𝒉 = β„Ž ([βˆ’cos 𝛽 sin 𝛾𝑖

cos 𝛾1

sin 𝛽 sin 𝛾1

] + [

βˆ’ cos 𝛾2

βˆ’ cos 𝛽 sin 𝛾2

sin 𝛽 sin 𝛾2

] + [cos 𝛽 sin 𝛾3

βˆ’ cos 𝛾3

sin 𝛽 sin 𝛾3

] + [

cos 𝛾4

cos 𝛽 sin 𝛾4

sin 𝛽 sin 𝛾4

])

(7)

(8)

(9)

Page 29: N McNamara - espace.library.uq.edu.au

N McNamara

22

= β„Ž ([

βˆ’cos 𝛽 sin 𝛾𝑖 βˆ’ cos 𝛾2 + cos 𝛽 sin 𝛾3 + cos 𝛾4

cos 𝛾1 βˆ’ cos 𝛽 sin 𝛾2 βˆ’ cos 𝛾3 + cos 𝛽 sin 𝛾4

sin 𝛽 sin 𝛾1 + sin 𝛽 sin 𝛾2 + sin 𝛽 sin 𝛾3 + sin 𝛽 sin 𝛾4

])

Combining Equations 3, 6 and 8 the equation of motion of the system is shown to be;

(𝑱�̇� + οΏ½Μ‡οΏ½) + 𝝎 Γ—(π‘±πŽ + 𝒉) = 𝑻EXT

Equation 11 is then represented as two simultaneous equations relating to the system and the

CMG control torque separately. Introducing the control torque vector term, 𝝉;

𝑱�̇� + 𝝎 Γ— π‘±πŽ = 𝝉 + 𝑻EXT

οΏ½Μ‡οΏ½ + 𝝎 Γ— 𝒉 = βˆ’π‰

For a desired attitude manoeuvre the required control torque, 𝝉, can be found. Rearranging

Equation 13 further, the non-linear equation describing the control system dynamics, referred

to as the torque equation, is;

οΏ½Μ‡οΏ½ = βˆ’π‰ βˆ’ 𝝎 Γ— 𝒉

5.2. Pseudoinverse and Minimum Norm Solution

The most rudimentary steering law is that of the pseudoinverse solution. By inverting the

torque equation (Equation 14) the required gimbal angle rates for a desired system angular

momentum state can be calculated. The time derivative of the CMG angular momentum

vector, 𝒉, is found by taking the derivative of Equation 9 such that;

οΏ½Μ‡οΏ½ = βˆ‘ 𝒉𝑖̇ (𝛾𝑖)

𝑁=4

𝑖=1

= 𝑨�̇�

where 𝜸 = (𝛾1, 𝛾2, 𝛾3, 𝛾4) is the gimbal angle vector, and 𝐴 is shown to be;

(10)

(11)

(12)

(13)

(14)

(15)

Page 30: N McNamara - espace.library.uq.edu.au

N McNamara

23

𝐴 = [βˆ’cos 𝛽 cos 𝛾𝑖

βˆ’sin 𝛾1

sin 𝛽 cos 𝛾1

sin 𝛾2

βˆ’ cos 𝛽 cos 𝛾2

sin 𝛽 cos 𝛾2

cos 𝛽 cos 𝛾3

sin 𝛾3

sin 𝛽 cos 𝛾3

βˆ’ sin 𝛾4

cos 𝛽 cos 𝛾4

sin 𝛽 cos 𝛾4

]

Such that;

οΏ½Μ‡οΏ½ = [βˆ’cos 𝛽 cos 𝛾𝑖

βˆ’sin 𝛾1

sin 𝛽 cos 𝛾1

sin 𝛾2

βˆ’ cos 𝛽 cos 𝛾2

sin 𝛽 cos 𝛾2

cos 𝛽 cos 𝛾3

sin 𝛾3

sin 𝛽 cos 𝛾3

βˆ’ sin 𝛾4

cos 𝛽 cos 𝛾4

sin 𝛽 cos 𝛾4

] [

𝛾1Μ‡

𝛾2Μ‡

𝛾3Μ‡

𝛾4Μ‡

]

Therefore, adding Equations 14 and 15 and letting οΏ½Μ‡οΏ½ ≑ 𝒖, the pseudoinverse solution is shown

to be;

𝑨�̇� = 𝒖 ≑ βˆ’π‰ βˆ’ 𝝎 Γ— 𝒉

∴ οΏ½Μ‡οΏ½ = 𝑨𝑇(𝑨𝑨𝑇)βˆ’1𝒖

Solving Equation 19 produces the minimum 2-norm vector of gimbal rates, οΏ½Μ‡οΏ½, that will result

in the desired attitude change. This solution drives the system towards a singularity state

which will cause the system to reach a singularity state. This can be viewed in terms of energy

to better understand why the system converges to a singular state.

The most energy efficient way to achieve an attitude change is to use angular momentum

contribution from all available sources. If the individual CMGs attribute equal angular

momentum changes to the overall system, the CMGs inherently approach a singular state

where β„Ž1 = Β±β„Ž2. As explained by Paradiso, 1991, β€œthe minimum norm property of the

pseudoinverse encourages the formation of singular states”. The pseudoinverse is only a

particular solution to the torque equation, other steering laws either use different approaches

to solve the equation or use the pseudoinverse solution and apply some form of modifying

scalar.

(16)

(17)

(18)

(19)

Page 31: N McNamara - espace.library.uq.edu.au

N McNamara

24

5.3. Momentum Envelope and CMG Gain

The momentum envelope of a given system represents the potential torque output of the

system in all planes of operation. More broadly it is used as a representation of the potential

total angular momentum for a given system along any given axis (Paradiso, 1991). It is a 3D

surface that represents the saturation limits of a system and any other internal singularity

states that arise. Previous findings from the literature detail the effects that different

configurations of CMG systems on the momentum envelope. Referring to Figure 10, the angle

of the pyramid sides, 𝛽, affects the overall shape of the outer surface of the system’s

momentum envelope. For the conceptual pyramid system, setting 𝛽 = 54.73Β° results in a

nearly spherical momentum envelope outer surface (Yoon & Tsiotas, 2002).

For larger satellite systems on orbital paths, a non-circular envelope may be employed to

maximise the potential angular momentum in a certain axis at the expense of another. For the

purposes of this project a spherical momentum envelope has been deemed the most relevant

when considering the practical applications of a spacesuit mission. Figure 11 depicts the

momentum envelope of the conceptual system.

Mathematically singularity states can be defined by the β€˜CMG gain’ of the system in the same

manner as Paradiso, 1991. Using the Jacobian matrix of the system as defined in Equation 5,

the CMG gain is defined as;

𝑀 = √|𝑨𝑨𝑇|

where 𝐽 = 𝐴 ∡ 𝐼𝐡 = 𝐴𝑔 = 0 for the simplified conceptual system case. As 𝑀 approaches 0 the

system approaches a singular state and loses control in one or more dimensions.

(20)

Page 32: N McNamara - espace.library.uq.edu.au

N McNamara

25

Figure 11 - Momentum envelope of conceptual system (Source: Yoon, 2004)

5.4. Elliptical and Parabolic Singularities

Before the steering laws can be addressed further clarification on the required control of the

system must be made. The singularity states can be categorised into two definitions;

β€˜hyperbolic’ and β€˜parabolic’. Using Figures 12 through 16 as reference, hyperbolic states are

escapable through null motion, on the momentum envelope they are represented by all

internal surfaces. Parabolic states are inescapable and are represented by the outer surfaces.

They occur when all CMGs have aligned angular momentum vectors. If the system enters a

parabolic singularity state then it can only be escaped by an external torque being applied.

The CMG steering law problem is complex due to the different behaviours required for

Page 33: N McNamara - espace.library.uq.edu.au

N McNamara

26

hyperbolic and parabolic states. Depending on the initial gimbal angles at the start of an

attitude manoeuvre, the system may need to transit through a hyperbolic state less the

system be restricted in operation for certain attitudes. In contrast, parabolic states must be

avoided at all costs.

Figure 12 - Hyperbolic singularity momentum envelope surface (Source: Yoon, 2004)

Figure 13 - Hyperbolic singularity CMG arrangement

Page 34: N McNamara - espace.library.uq.edu.au

N McNamara

27

Figure 14 – Parabolic singularity momentum envelope surface (Source: Yoon, 2004)

Figure 15 - Parabolic singularity CMG arrangement

Page 35: N McNamara - espace.library.uq.edu.au

N McNamara

28

Figure 16 - Momentum envelope of pyramid configuration CMG system (Source: Yoon, 2004)

Page 36: N McNamara - espace.library.uq.edu.au

N McNamara

29

6. Steering Laws

There are many proposed steering laws that can be found in the literature that have potential

to control a spacesuit CMG system and avoid the described singularity states. Viewing the

steering laws in a broad sense there are two distinct types; singularity-avoidance laws and

singularity-robust laws (Jones, Zeledon, & Peck, 2012). The latter allows the system to enter a

hyperbolic singularity state (escapable and denoted by the internal surface on the momentum

envelope), in some circumstances where the transit through the singularity can be controlled.

Singularity-robust laws are the more modern method of control of the two types. The former,

singularity-avoidance laws, do not allow the system to enter a singularity state in any

circumstance. Neither of the steering law types allow the system to enter a parabolic,

inescapable singularity state. The only way to restore control to a saturated system in this

state is to apply an external torque which, for obvious, reasons is outside the control of any

potential steering law (Gui & Vukovich, 2016).

Further to this most modern steering laws can be placed into one of six categories that

differentiate between distinctive design mentalities (Kurokawa, 2007);

1. Moore-Penrose Pseudoinverse

2. Offline Planning

3. Preferred Gimbal Angle/Gimbal Reorientation

4. Linearly Constrained

5. Gradient Weighting/Null Motion

6. Singularity Robust Inverse

No singularity control

Singularity-avoidance

Singularity-robust

Page 37: N McNamara - espace.library.uq.edu.au

N McNamara

30

The listed design mentality categories have different advantages and characteristics. Although

each may be broken down into further subcategories these are mostly trivial and do not

change the major design characteristics of the various solutions (Jones, Zeledon, & Peck,

2012). The special case of the Moore-Penrose Pseudoinverse has no inherent protocol to avoid

or deal with singularities but it has been used in scenarios where the potential momentum

envelope is significantly greater than the required envelope for a given mission/application. It

is worth noting that most steering laws use the Moore-Penrose solution as a basis and apply

various scaling factors to control singularities (NASA TM, 1972). For this reason, although it is

not suitable to use this law with a spacesuit system, it has been analysed to gain further insight

into the other steering laws.

The investigated control laws have been assessed in several areas as they pertain to use with

a spacesuit. The points of qualitative assessment are;

β€’ Singularity handling method,

β€’ Induced error and perturbations,

β€’ System efficiency and the full use of the hardware momentum envelope,

β€’ Convergence to singularity arrangements, and

β€’ Potential for generalization to other CMG arrangements.

The last assessment criterium has been included as it is a good indication of how the steering

law will cope with partial system failure. If for reasons outside the control of the CMG

controller one or more CMGs fail, the controller must be adaptable to configurations other

than the nominally designed case.

Page 38: N McNamara - espace.library.uq.edu.au

N McNamara

31

6.1. No Singularity Avoidance

6.1.1. Moore-Penrose Pseudoinverse

As previously discussed in Section 5.2, the Moore=Penrose Pseudoinverse (MPI) solution is

the basic form of the steering logic and is the solution to Equation 18 (Jung & Tsiotras, 2008);

�̇�𝑁 = 𝑨𝑇(𝑨𝑨𝑇)βˆ’1𝒖

The MPI solution is one that minimizes the required system input, gimbal angle rate (�̇�𝑁), to

achieve the desired system output, total CMG torqueing (𝒖). Jung & Tsiotras (2008) define the

MPI solution as the one that minimizes the input gimbal rates;

minοΏ½Μ‡οΏ½

||οΏ½Μ‡οΏ½||2

subject to 𝑨�̇� = 𝒖

If a system employs an MPI steering controller then the system will eventually converge to a

singular state without fail (Jung & Tsiotras, 2008).

Gui & Vukovich (2016) conducted an analysis on potential steering laws for a spacecraft with

two parallel CMGs. The authors found that a controller employing an unmodified MPI solution

results in satisfactory performance up until a saturation limit was reached. At this point the

system was unable to escape the singularity. Further work by the George C. Marshall Flight

Center, NASA (1972) found that to make an unmodified MPI controller viable, the mission had

to be restricted to 50-60% of the potential momentum envelope cap abilities of the CMG

system. The authors also noted that the unmodified MPI controller can control the system in

the instance of a single CMG failure for a 4-CMG system. At the point of failure, the system

was still capable of calculating the required gimbal rates to achieve the desired attitude

change (NASA TM, 1972).

(21)

Page 39: N McNamara - espace.library.uq.edu.au

N McNamara

32

6.2. Singularity-Avoidance

6.2.1. Offline Planning

Offline planning is an all-encompassing term and relates to any steering law logic which uses

predefined data sets to control the CMG system for a given scenario. For instance, a system

may be modelled for a range of potential attitude changes. If any particular manoeuvres

introduce singularity states to the system then these can be adjusted. For any similar

manoeuvers in the future the system can use memory based logic to avoid the singularity

state.

Significant research has been done in this area by Paradiso (1991). The author investigated

the use of heuristic optimization of pre-planned gimbal angle inputs for a range of torque-

producing trajectories. After the database of planned manoeuvres has been established then

a search protocol must be implemented. The search protocol is responsible for analysing the

system state and then comparing to cases stored in the database. Although this methodology

has the potential for optimization it does not allow for adaption to unplanned system changes

or states. The ISS has employed offline planning for the control of its 6 CMG control system

but this is highly dependent on the fact that the station only carried out schedule attitude

changes (Gurrisi, et al., 2010).

For use with a spacesuit, offline planning steering control is likely unfeasible. Unlike the ISS, a

user driven spacesuit must be subjectable to range to a range of attitude manoeuvers. Nor is

the steering law capable of dealing with system changes or component failure. Although it is

possible to model such a system accurately given enough computational power, such

resources are likely unavailable onboard a spacesuit system.

Page 40: N McNamara - espace.library.uq.edu.au

N McNamara

33

6.2.2. Preferred Gimbal Angle/Gimbal Reorientation

An extension to offline planning is the preferred gimbal angle method. First proposed by

Vadali, Oh & Walker (1990), the method uses the MPI solution to find the most efficient initial

gimbal angles for a set of required attitude manoeuvres. The system is modelled for initial

states near saturation (the outer momentum envelope surface) and back integration using the

MPI solution is used to determine all attitude changes that are possible without the system

entering a singularity state, either hyperbolic or parabolic. The results are then compared to

the potential mission profile and then the most appropriate initial gimbal angle position is

determined (Jones, Zeledon, & Peck, 2012).

The preferred gimbal angle controller has the same advantages and characteristic as the

offline planning method. Vadali, Oh & Walker (1990) conducted numerical simulations of their

proposed steering logic on the Soviet space station MIR, and Space Station Freedom (the

original concept mission that later became the ISS mission).

The authors found that optimizing the initial gimbal angles allowed the system to utilize

regions of the momentum envelope closer to singularity states than the offline planning

method at the cost of full use of the rest of the envelope (Vadali, Oh, & Walker, 1990). Like

offline planning, the preferred gimbal angle method is still highly dependent on the foresight

or potential future attitude manoeuvers and is susceptible to unforeseen system changes and

failure.

Page 41: N McNamara - espace.library.uq.edu.au

N McNamara

34

6.2.3. Linearly Constrained

The final singularity-avoidance law is the linearly constrained method. The steering law first

proposed by Kurokawa (1997), is a simple geometric restraint placed on the relative gimbal

angles. The author present the simple case for a 4 CMG pyramid configuration which is very

close to the conceptual system proposed in this report. Noting that internal singularity states

occur at any point where the angular momentum vector of one CMG aligns or negates that of

another, ie;

π’‰π’Š = ±𝒉𝑗

By restricting the relative angles between the gimbals then singularity states along one or

more specific axes can be avoided. Using the case presented by Kurokawa (1997) the system

can be made non-singular by applying the constraint function;

πœƒ1 βˆ’ πœƒ2 + πœƒ3 βˆ’ πœƒ4 = 0

A geometric representation of the constraint equation for a system of 3 CMGs can be seen in

Figure 17. A draw back of the linearly constrained steering law is significant reduction in the

useable momentum envelope. Figure 18 shows the restricted momentum envelope

workspace for a 4 CMG pyramid system that prevents singularities in the 𝑧-axis, note the

shaded region is the lost momentum envelope as a result of the linearly constrained steering

law.

(22)

Page 42: N McNamara - espace.library.uq.edu.au

N McNamara

35

Figure 17 – Constrained steering law applied to system of 3 CMGs (Source: Jones, et al., 2012)

Figure 18 - Effects of constrained steering law on momentum envelope (Source: Kurokawa, 1997)

Page 43: N McNamara - espace.library.uq.edu.au

N McNamara

36

6.3. Singularity-Robust

6.3.1. Gradient/Null Motion Weighted Matrix

The first type of singularity-robust steering law allows for the system to transit through

internal singularities by utilizing null motion. Null motion is the process of actuating gimbals

in such a way as to produce a net zero torque on the overall system, but which results in a

reorientation of the gimballed CMGs (Bedrossian, Paradiso, & Bergmann, 1990).

Mathematically the null motion method uses the CMG gain value, 𝑀 = βˆšπ‘¨π‘¨π‘‡, and weighting

matrix, π‘˜[𝑰], to determine the degree of introduced null motion in the system. Referring to

general MPI solution of Equation 18;

minοΏ½Μ‡οΏ½

||οΏ½Μ‡οΏ½||2

subject to 𝑨�̇� = 𝒖

The weight null motion solution has the form;

οΏ½Μ‡οΏ½ = οΏ½Μ‡οΏ½MPI + π‘˜[𝑰] βˆ™ οΏ½Μ‡οΏ½NM

where, οΏ½Μ‡οΏ½MPI is the general gimbal angle solution in line with the MPI method and οΏ½Μ‡οΏ½NM is the

introduced null motion factor. There is an inversely proportional relation between CMG gain,

𝑀, and the weighted matrix, π‘˜[𝑰];

𝑀 = βˆšπ‘¨π‘¨π‘‡ ∝ π‘˜[𝑰]βˆ’1

Other sources from the literature detail alternative implementations of the null motion

methodology. Systems can employ step functions to introduce null motion into the gimbal

rate solutions (Yoshikawa, 1977), pair the null motion with scissoring limit laws that link

partner CMGs (Bedrossian, Paradiso, & Bergmann, 1990), or use null motion principles to

create oscillations about a singularity state to maximise the useable work envelope (Hefner &

McKenzie, 1983).

(23)

(24)

Page 44: N McNamara - espace.library.uq.edu.au

N McNamara

37

Other works by Jin & Hwang (2011) and Zhang & Fang (2013) investigate the potential of back-

stepping control methods for use with double-gimbal VSCMG systems. The authors present a

dual-loop control system comprising of an attitude loop and an angular speed loop. Although

the authors investigate the control of a variable speed CMG system which is out of scope for

this project, the null motion principles are applicable to a constant speed CMG system. Zhang

& Fang (2013) present the control law using Euler notation, Jin & Hwang (2011) develop the

controller using quaternion notation however both follow the same logic (Jin & Hwang, 2011;

Zhang & Fang, 2013). As the system approaches the singularity state a small random

perturbation in null motion is introduced into the attitude control loop to allow the spacecraft

to reach the desired attitude while maintaining a non-singular gyroscope orientation (Zhang

& Fang, 2013).

Compared to singularity-avoidance laws, the useable work envelope available for a system

using a null motion control is greater. Although some methods introduced random null motion

perturbations rather than strictly controlled rates, both methods allow the system to reach

near singularity points, and oscillate about and through them. For a smaller system on a

spacesuit where hardware sizing is limited, the full utilization of the work envelope is crucial.

6.3.2. Singularity-Robust Inverse

The final steering law method, and the most modern, is an extension on the null motion

principles. Most null motion controls intentionally add error to the solution to oscillate the

system around and/or through singularities. Although easily implemented while still creating

useable torque from respective CMGs near singularities, the system torque is only accurate in

a small region of the momentum envelope. The closer to a singularity the system gets, the

Page 45: N McNamara - espace.library.uq.edu.au

N McNamara

38

greater the attitude error. There is trade-off between the useable work envelope and accuracy

of attitude at different points in the envelope.

If random perturbations are introduced into the system then there is also a risk of control

gimbal rates that exceed hardware limitations. There is significant research on the topic of

singularity-robust inverse (SRI) steering laws that can be found in the literature.

Wie, Bailey & Heiberg (2001) present the generalized SRI solution. In comparison to the null

motion solution, the SRI solution introduces a new term into Equation 21 in an attempt to

calculate gimbal rates that produce an accurate, non-singular result. Taking Equation 21 and

adding the additional term, πœŒπ’, the general SRI solution is;

�̇�𝑁 = 𝑨𝑇(𝑨𝑨𝑇)βˆ’1𝒖 + πœŒπ’

where 𝒏 is the 1Γ—4 matrix of form [𝐢1, 𝐢2, 𝐢3𝐢4] containing the optimized singularity measure

values that determine the amount of null motion from each CMG, and 𝜌 is the step function;

𝜌 = { 𝑀6 for 𝑀 β‰₯ 1π‘€βˆ’6 for 𝑀 < 1

Note, 𝑀 is the CMG gain of the system as defined previously.

The benefit of this solution is that null motion introduced into the system is proportional to

the state of the system and can be accounted for. This method is superior to the random

perturbations introduced with other methods and reduces inherent error with attitude

changes near singularities.

Jung & Tsiotras (2008) investigated the potential to add further terms to Equation 25. The

authors tested various SRI solutions on a 3 CMG system and concluded that the induced

torque error was reduced and that the energy efficiency of the system was greater when

compared to an MPI solution with randomly induced torqueing when near singularities.

(25)

Page 46: N McNamara - espace.library.uq.edu.au

N McNamara

39

6.4. Qualitative Summary

Table 2 summarises the properties of the detailed steering laws against the metrics listed in

Section 6.0.

Table 2 - Qualitative comparison summary

Although each steering law methodology has various advantages and characteristics which

are applicable to a spacesuit system, the singularity-robust inverse method is the best fit for

purpose. The full use of the momentum envelope with a smaller package system like a

spacesuit is crucial. It is also essential that the system can adapt to any hardware changes or

error, system failure with a user outside of the parent spacecraft is unacceptable. The SRI

solution provides the user with redundancy without the need for excessive computational

Singularity Handling Method

Induced Error

Momentum Envelope Efficiency

Convergence to Singularity Arrangements

Generalization Capabilities for System Changes

Moore-Penrose Pseudoinverse

None None Full use Possible Fully adaptable

Offline Planning

Optimized pre-modelling None Slightly reduced Impossible Requires extensive computational power

Preferred Gimbal Angle

Optimized initial gimbal orientations

None Slightly reduced Impossible Requires extensive computational power and system adjustment

Linearly Constrained

Restricted relative gimbal angles

None Reduced dependent on initial gimbal angles

Impossible Fully adaptable

Null Motion Randomly induced torque scaling with proximity to singularities

Large near singularity

Full use but with large induced error in segments

Possible Fully adaptable

Singularity-Robust Inverse

Optimized null motion scaling factor from system Jacobian inertial matrix

Controlled near singularity

Full use Possible Fully adaptable

Page 47: N McNamara - espace.library.uq.edu.au

N McNamara

40

power and the full use of the momentum envelope available unlike singularity-avoidance

methods.

Although the weighted null motion solution is also feasible, the reduced error with the SRI

solution is beneficial. If missions call for finer motor control of instruments for maintenance

or scientific processes, the reduced error results in greater attitude stabilization.

7. Simulation and Results

The SRI solution was tested against the MPI solution for the conceptual system. Figure 10

details the geometry of the system. The simulation was run using a time-step interval of 0.001

seconds. The work of Yime-Rodriguez et al. (2014) has been used heavily in the construction

of the computational model. Several simplifications have been made; the centre of mass of

the system has been placed at the base of the pyramid as indicated. The initial gimbal angles

have been set to 0Β° for all gimbals, and therefore the initial angular momentum vector of the

system is also 0 such that π’‰πŸŽ = [𝟎, 𝟎, 𝟎]𝑻. The pyramid skew angle has been set at 53.47Β° to

ensure a nearly spherical momentum envelope. The angular momentum magnitude for all

flywheels has been set to a nominal value of 1 π‘˜π‘” βˆ™ π‘š/𝑠2 to ensure uniformity and to simplify

the model response. A gimbal rate limit of Β±25Β°/𝑠 and a gimbal angle limit of Β±100Β° was used

as per consensus from the literature on feasible hardware limits (Jung & Tsiotras, 2008).

The model requests a constant torque demand from the system in the π‘₯-axis (roll-axis) such

that οΏ½Μ‡οΏ½ ≑ 𝒖 = [𝟏, 𝟎, 𝟎]𝑻. Applying the Euler method, at each time step the model calculates

the gimbal rates using the MPI or SRI steering law solution for each case respectively. An

estimate of the angular momentum vector, 𝒉, is then found. The singularity state of the CMG

Page 48: N McNamara - espace.library.uq.edu.au

N McNamara

41

configuration can then be visualized by plotting the CMG gain against the magnitude of

angular momentum in the roll-axis, ie 𝑀 versus |β„Žπ‘₯|.

For the MPI test, the minimum gimbal rate solution to the identity equation οΏ½Μ‡οΏ½ = 𝑨𝑇(𝑨𝑨𝑇)βˆ’1𝒖

was found for the final system 𝐴 matrix;

𝐴 = [βˆ’cos 𝛽 cos 𝛾𝑖

βˆ’sin 𝛾1

sin 𝛽 cos 𝛾1

sin 𝛾2

βˆ’ cos 𝛽 cos 𝛾2

sin 𝛽 cos 𝛾2

cos 𝛽 cos 𝛾3

sin 𝛾3

sin 𝛽 cos 𝛾3

βˆ’ sin 𝛾4

cos 𝛽 cos 𝛾4

sin 𝛽 cos 𝛾4

]

For the SRI test the same torque demand was used. The additional null motion component

was added to the minimum solution equation such that οΏ½Μ‡οΏ½ = 𝑨𝑇(𝑨𝑨𝑇 + πœŒπ‘°)βˆ’1𝒖. The scaling

factor, 𝜌, was set using the values found from experimental optimization by Jung & Tsiotras

(2008) such that;

IF 𝑀 > 𝑀critical;

𝜌 = 0

ELSE

IF 𝜌0

𝑀< 𝜌MAX;

𝜌 =𝜌0

𝑀

ELSE

𝜌 = 𝜌MAX

where 𝑀critical = 1.0, 𝜌0 = 0.1, and 𝜌MAX = 0.2. This ensures that the introduced null motion

in the SRI solution is low when the system is far from a singular state, and increases as the

system approaches a singular state. Note, the system is not singular at 𝑀 = 𝑀critical. The

Page 49: N McNamara - espace.library.uq.edu.au

N McNamara

42

induced torque error of the SRI system can be calculated by computing the difference

between the input angular momentum command, οΏ½Μ‡οΏ½ ≑ 𝒖 = [𝟏, 𝟎, 𝟎]𝑻̇ , and the angular

momentum that would result from the calculated gimbal rates at each time step, ie;

οΏ½Μ‡οΏ½error = |οΏ½Μ‡οΏ½input βˆ’ (𝑨 βˆ™ οΏ½Μ‡οΏ½calculated)|

Figures 19 and 20 show the results. The raw data points are given in Appendix B.

Figure 19 - Comparison of MPI and SRI Solutions

Figure 20 - SRI Solution Induced Torque Error from Input Command

(26)

Page 50: N McNamara - espace.library.uq.edu.au

N McNamara

43

8. Discussion

As expected the MPI system rapidly approaches a singular state at 𝒉π‘₯ β‰… 1.15 corresponding

to the angular momentum alignment of CMGs 1 and 3 at 𝜸 = [βˆ’πŸ—πŸŽ, 𝟎, πŸ—πŸŽ, 𝟎]𝑻 (Β°). This

position corresponds to an inescapable singularity state (Figures 14 and 15). The parabolic

response of the CMG gain near the singularity can be seen in Figure 19. At this point, CMGs 2

and 4 are not capable of producing a change in angular momentum in the roll-axis and

therefore the system rapidly approaches the singularity state.

Using the SRI system, it is shown that once the critical gain limit is reached then the control

law introduces a disturbing factor to avoid the singularity. As shown in Figure 20, the system

shows a deflection in the roll (π‘₯), pitch (𝑦) and yaw (𝑧) axes near the singularity state. Although

it is still not possible to pass directly through the singularity with no introduced error, the

system can successfully reach an angular momentum vector very-near-to while maintaining a

higher CMG gain.

The results of the conceptual system model support the findings from the qualitative

comparison of the steering laws. The singularity-robust inverse solution avoided the

singularity state by inducing a small, controlled torqueing error when near a singularity state.

For implementation with a spacesuit the SRI solution should be used over an MPI solution. It

does not require excessive computational power like other control methods and it effectively

avoids singularity states. The trade-off of angular momentum error is deemed acceptable

when compared to the confidence that the system will avoid a singularity. In a worst-case

scenario for an astronaut during an EVA, if a singularity state is entered and there is no foreign

body with induce an outside torque on the system, then the astronaut may be stuck.

Further consideration should be made before implementing a CMG system into a spacesuit.

Although current computational power precludes aforementioned steering methods at the

Page 51: N McNamara - espace.library.uq.edu.au

N McNamara

44

current time, faster computers or quantum computers may allow such methods to be used in

the future. The essential factor as to when machine learning and search algorithm dependent

laws become available is the ability to recalculate nominal gimbal rates in the event of a

system change or component failure in a timely manner.

Alternatively, different hardware configurations other than that detailed in this report can be

investigated. Different configurations other than a pyramid can be used to alter the

momentum envelope of the spacesuit. Double-gimbal CMG systems are possible where a

single flywheel can be actuated as to produce a torque vector in any direction in 3D-space. In

theory, a single double-gimbal CMG could be used to control a spacesuit but there is significant

strain placed on the gimbal mounting structure as a result of transverse bending during gimbal

actuation (Zhang & Fang, 2013).

Additionally, there is significant knowledge available in the literature regarding variable-speed

CMGs which allow for the flywheel speed of each CMG to be altered independently. Using the

example from the simulation in Section XXX, the singularity state could be avoided by altering

the speed of flywheels 1 and 3 in such a way that would allow the system to reach an angular

momentum vector of 𝒉 = [1.15, 0, 0]𝑇 with a CMG gain, 𝑀, > 0 (Jin & Hwang, 2011). A new

area of research emerging is the potential to use a variable-speed CMG system for attitude

control and power storage. High inertia flywheels have been used on satellite systems as a

form of energy storage for some time. Yoon (2004) provides an overview of spacecraft control

and power management using such a system. Although feasible and potentially viable the

author notes the additional electrical motor and linkages required to enable the braking and

acceleration of the flywheel. Although conceivable for a satellite or spacecraft sized vehicle

the additional mass of the electrical motor as a percentage of total mass is much higher for a

spacesuit.

Page 52: N McNamara - espace.library.uq.edu.au

N McNamara

45

9. Conclusion

The project has met the initial aim of conducting a qualitative and quantitative comparison of

control moment gyroscope system control laws for use with a spacesuit sized vehicle. Control

moment gyroscopes are a viable technology for use with a spacesuit. Utilizing momentum

exchange principles, a spacesuit could conserve resources and energy as compared to the

modern day manned-manoeuvring units. Providing astronauts with a means of attitude

control and stabilization will lead to the reduction of EVA related injuries of the hands and

forearms. It will also provide a solid foundation from which astronauts will be able to conduct

maintenance and scientific work.

Singularity-avoidance and singularity-robust control laws which are currently used with

satellite and larger-scale spacecraft can potentially be adapted for use with spacesuit. Broadly

speaking any given control law either induces a controlled torqueing error when near

singularities or restricts the available momentum envelope to avoid singularities altogether.

The singularity-robust inverse control law is deemed the most viable option. As shown in the

simulation results, a proportional torqueing error is introduced when the CMG gain of the

system approaches 0. The control law successfully navigates around singularities. This method

allows for the full use of the potential momentum envelope for the given hardware. Compared

to a larger spacecraft the mass of the CMGs as a percentage of total mass is much higher for

a spacesuit sized system. Maintaining full use of the momentum envelope is crucial. The

control law is also capable of accounting for system changes and/or component failure

without the need for excessive computational modelling or hardware adjustment.

Implementing a CMG control system, utilizing the singularity-robust inverse control law, into

a spacesuit will allow future astronauts to navigate space more efficiently than what is

possible today.

Page 53: N McNamara - espace.library.uq.edu.au

N McNamara

46

10. References

Bedrossian, N., Paradiso, J., & Bergmann, E. (1990). Steering Law Design for Redundant Single-

Gimbal Control Moment Gyroscopes. Journal of Guidance, Control and Dynamics,

1083-1089.

Brown, D., & Peck, M. (2009). Energetics of Control Moment Gyroscopes as Joint Actuators.

Journal of Guidance, Control, and Dynamics, 1871-1883.

Carpenter, M., Jackson, K., Cohanim, B., Duda, K. R., Rize, J., Dopart, C., . . . Zumbado, J. R.

(2013). Next Generation Maneuvering System with CMGs for EVA Near Low-Gravity

Objects. Houston: American Institude of Aeronautics and Astronautics.

Cheng, J. P. (2010). Dynamic Model of an Astronaut Equipped with a Manned Maneuvering

Unit in Virtual Reality. Journal of Aerospace Engineering, 139-145.

Danaher, P., Tanaka, K., & Hargens, A. R. (2005, April 4). Mechanical Counter-Pressure vs. Gas-

Pressurized Spacesuit Gloves: Grip and Sensitivity. Aviation, Space, and Environmental

Medicine, pp. 381-384.

Dawson, R. (2013). Orbits and Locked Gimbals. New York: Springer Science + Business Media.

Gui, H., & Vukovich, G. (2016). Attitude Stabilization of a Spacecraft with Two Parallel Control

Moment Gyroscopes. Journal of Guidance, Control and Dynamics, 724-731.

Gurrisi, C., Seidel, R., Dickerson, S., Didziulis, S., Frantz, P., & Ferguson, K. (2010). Space Station

Control Moment Gyroscope Lessons Learned. Houston: NASA.

Harland, D. M. (2010). NASA's Moon Program - Paving the Way for Apollo 11. Dordrecht:

Springer.

Hefner, R., & McKenzie, C. (1983). A Technique for Maximizing the Torque Capability of

Control Moment Gyroscopes. Proceedings of the AAS/AIAA Astrodynamics Conference

(pp. 905-920). Lake Placid: AAS.

Jennings, R., & Bagian, J. (1996). Musculoskeletal injuries review in the U.S. space program.

Aviation, Space, and Environmental Medicine, pp. 762-766.

Jin, J., & Hwang, I. (2011). Attitude Control of a Spacecraft with Single Variable-Speed Control

Moment Gyroscope. Journal of Guidance, Control and Dynamics, 1920-1924.

Jones, L. L., Zeledon, R. A., & Peck, M. A. (2012). Generalized Framework for Linearly

Constrained Control Moment Fyro Steering. Journal of Guidance, Control and

Dynamics, 1094-1103.

Jung, D., & Tsiotras, P. (2008). An Experimental Comparison of CMG Steering Control Laws.

Atlanta: Georgia Institute of Technology.

Page 54: N McNamara - espace.library.uq.edu.au

N McNamara

47

Kim, D. (2011). A Novel Integrated Spacecraft Attitude Control System Using Variable Speed

Control Moment Gyroscopes: A Lyuapunov-Based Approach. University of Florida.

Kurokawa, H. (2007). Survey of Theory and Steering Laws of Single-Gimbal Control Moment

Gyroscopes. Journal of Guidance, Control and Dynamics, 1331-1340.

Leve, F. A., Hamilton, B. J., & Peck, M. A. (2015). Spacecraft Momentum Control Systems. New

York: Springer International.

MacKunis, W., Dupree, K., Fitz-Coy, N., & Dixon, W. (2008). Adaptive Satellite Attitude Control

in the Presence of Inertia and CMG Gimbal Friction Uncertainties. The Journal of

Astronautical Sciences, 121-134.

McMahon, J., & Schaub, H. (2009). Simplified Singularity Avoidance Using Variable-Speed

Control Moment Gyroscope Null Motion. Boulder: Journal of Guidance, Control and

Dynamics.

NASA TM. (1972). A Comparison of CMG Steering Laws for High Energy Astronomy

Observatories. NASA Marshall Space Flight Center.

Opperman, R. A., Waldie, J. M., Natapoff, A., Newman, D. J., & Jones, J. A. (2010, October).

Probability of Spacesuit-Induced Fingernail Trauma Is Associated with Hand

Circumference. Aviation, Space, and Environmental Medicine, pp. 907-913.

Paradiso, J. (1991). Global Steering of Signle GImballed cOntrol Moment Gyrosopes Using a

Directed Search. AIAA Guidance and Control Conference. New Orleans: The Charles

Stark Draper Laboratory.

Russel, S. P., Spencer, V., & Metrocavage, K. (2008). On-Orbit Propulsion and Methods of

Momentum Management for the International Space Station. AIAA/ASME/SAE/ASEE

Joint Propulsion Conference & Exhibit. Hartford: JPL.

Tewari, A. (2011). Automatic Control of Atmospheric and Space Vehicles. London: Springer.

Thomas, K. S., & McMann, H. J. (2006). US Spacesuits. Chichester: Springer Science and

Business Media.

Vadali, S., Oh, H., & Walker, S. (1990). Preferred Gimbal Angles for Single Gimbal Control

Moment Gyroscope. Journal of Guidance, Control, and Dynamics, 1090-1095.

Votel, R., & Sinclair, D. (2013). Comparison of Control Moment Gyros and Reaction Wheels for

Small Earth-Observing Satellites. Annual AIAA/USU Conference on Small Satellites.

Utah: Utah State University.

West, W., & Witt, V. (2010). Post-Shuttle EVA Operations on ISS. Houston, Texas: American

Institute of Aeronautics and Astronautics.

Wie, B. (2001). Singularity Robust Steering Logic for Redundant Single-Gimbal Control

Moment Gyroscopes. Journal of Guidance, Control and Dynamics, 865-872.

Page 55: N McNamara - espace.library.uq.edu.au

N McNamara

48

Yime-Rodriguez, E., Pena-Cortes, C., & Rojas-Contreras, M. (2014). The Dynamic Model of a

Four Control Moment Gyroscope System. Colombia: SciELO.

Yoon, H., & Tsiotas, P. (2002). Spacecraft Adaptive Attitude and Power Tracking with Variable

Speed Control Moment Gyroscopes. Journal of Guidance, Control and Dynamics, 1081-

1090.

Yoshikawa, T. (1977). Steering Law for Roof Type Configuration Control Moment Gyro System.

Automatica, 359-368.

Zhang, H., & Fang, J. (2013). Robust Backstepping Control for Agile Satellite Using Double-

Gimbal Variable-Speed Control Moment Gyroscope. Journal of Guidance, Control and

Dynamics, 1356-1367.

Page 56: N McNamara - espace.library.uq.edu.au

N McNamara

49

11. Appendices

11.1. Appendix A – CMG Torque Derivation from First Principles

Brown & Peck (2009) outline the derivation of torque applied by a control moment gyroscope.

The angular momentum of a CMG about the centre of mass is the sum of the momentum of the

flywheel and the gimbal structure;

β„ŽCMG = 𝐼𝑔 βˆ™ πœ”πΊπ‘‚ + πΌπ‘Ÿ βˆ™ πœ”

𝑅𝑂

= 𝐼𝑔 βˆ™ (πœ”πΊπ΅ + πœ”

𝐡𝑂) + πΌπ‘Ÿ βˆ™ (πœ”

𝑅𝐺 + πœ”

𝐺𝐡 + πœ”

𝐡𝑂)

where 𝐼𝑔 and πΌπ‘Ÿ is the moment of inertia of the gimbal and rotor respectively, and πœ”π‘–

𝑗 is the angular

velocity of in the 𝑖 frame relative to the 𝑗 frame for; rotor frame, 𝑅, gimbal frame, 𝐺, body frame, 𝐡,

and origin frame, 𝑂.

Letting;

πΌπ‘Ÿ βˆ™ πœ”π‘…πΊ = β„Žπ‘Ÿ

∴ β„ŽCMG = 𝐼CMG βˆ™ (πœ”πΊπ΅ + πœ”

𝐡𝑂) + β„Žπ‘Ÿ

where, 𝐼𝑔 + πΌπ‘Ÿ = 𝐼CMG. Taking the time derivative in the origin frame;

𝜏CMG = βˆ’β„ŽCMG𝑂 βˆ’ πœπ‘”

where;

β„ŽCMG𝑂 = 𝐼CMG βˆ™ (οΏ½ΜˆοΏ½π‘”Γ—πœ”

𝐡(𝐡𝑂

)βˆ’ οΏ½Μ‡οΏ½π‘”Γ—πœ”

𝐡𝑂) + (�̇�𝑔 + πœ”

𝐡𝑂) Γ— (𝐼CMG βˆ™ (�̇�𝑔 + πœ”

𝐡𝑂) + β„Žπ‘Ÿ)

where;

�̇�𝑔 = πœ”πΊπ΅

Page 57: N McNamara - espace.library.uq.edu.au

N McNamara

50

For the case where a CMG has a spherical momentum envelope (as is the case with most constant

speed flywheel systems), [(�̇�𝑔 + πœ”π΅

𝑂) Γ— (𝐼CMG βˆ™ (�̇�𝑔 + πœ”π΅

𝑂) + β„Žπ‘Ÿ)] can be eliminated (Brown & Peck,

2009).

∴ β„ŽCMG𝑂 = 𝐼CMG βˆ™ (οΏ½ΜˆοΏ½π‘”Γ—πœ”

𝐡(𝐡𝑂

)βˆ’ οΏ½Μ‡οΏ½π‘”Γ—πœ”

𝐡𝑂)

This equation can then be split into two separate equations representing the torques applied by the

gimbal structure and torque transferred to the parent system.

Neglecting friction, electro-magnetic forces, and flexible body effects;

𝜏CMG = βˆ’οΏ½Μ‡οΏ½π‘”Γ—β„Žπ‘Ÿ

πœπ‘” = 𝐼CMG�̈� + 𝐼CMGπœ”π΅(

𝐡𝑂

)βˆ™ 𝑔 + (πœ”

π΅π‘‚Γ—β„Žπ‘Ÿ) βˆ™ 𝑔

Page 58: N McNamara - espace.library.uq.edu.au

N McNamara

51

11.2. Appendix B – Simulation Data

The data in this section is represented with an angular momentum interval of 0.01 from 0 to 2.25.

MPI SRI

𝒉𝒙 𝑀 𝑀 Ξ”πœπ‘₯ Ξ”πœπ‘¦ Ξ”πœπ‘§

0.00 0.0000 0.0000 0 0 0

0.01 1.0800 1.0800 0 0 0

0.02 1.0790 1.0790 0 0 0

0.03 1.0765 1.0765 0 0 0

0.04 1.0733 1.0733 0 0 0

0.05 1.0700 1.0700 0 0 0

0.06 1.0667 1.0667 0 0 0

0.07 1.0629 1.0629 0 0 0

0.08 1.0588 1.0588 0 0 0

0.09 1.0544 1.0544 0 0 0

0.10 1.0500 1.0500 0 0 0

0.11 1.0452 1.0452 0 0 0

0.12 1.0398 1.0398 0 0 0

0.13 1.0344 1.0344 0 0 0

0.14 1.0293 1.0293 0 0 0

0.15 1.0250 1.0250 0 0 0

0.16 1.0213 1.0213 0 0 0

0.17 1.0177 1.0177 0 0 0

0.18 1.0146 1.0146 0 0 0

0.19 1.0119 1.0119 0 0 0

0.20 1.0100 1.0100 0 0 0

0.21 1.0085 1.0085 0 0 0

0.22 1.0072 1.0072 0 0 0

0.23 1.0060 1.0060 0 0 0

0.24 1.0053 1.0053 0 0 0

0.25 1.0050 1.0050 0 0 0

0.26 1.0053 1.0053 0 0 0

0.27 1.0061 1.0061 0 0 0

0.28 1.0072 1.0072 0 0 0

0.29 1.0086 1.0086 0 0 0

0.30 1.0100 1.0100 0 0 0

0.31 1.0118 1.0118 0 0 0

0.32 1.0140 1.0140 0 0 0

0.33 1.0167 1.0167 0 0 0

0.34 1.0194 1.0194 0 0 0

0.35 1.0220 1.0220 0 0 0

0.36 1.0245 1.0245 0 0 0

0.37 1.0270 1.0270 0 0 0

Page 59: N McNamara - espace.library.uq.edu.au

N McNamara

52

0.38 1.0296 1.0296 0 0 0

0.39 1.0323 1.0323 0 0 0

0.40 1.0350 1.0350 0 0 0

0.41 1.0377 1.0377 0 0 0

0.42 1.0404 1.0404 0 0 0

0.43 1.0432 1.0432 0 0 0

0.44 1.0463 1.0463 0 0 0

0.45 1.0500 1.0500 0 0 0

0.46 1.0546 1.0546 0 0 0

0.47 1.0601 1.0601 0 0 0

0.48 1.0664 1.0664 0 0 0

0.49 1.0731 1.0731 0 0 0

0.50 1.0800 1.0800 0 0 0

0.51 1.0871 1.0871 0 0 0

0.52 1.0948 1.0948 0 0 0

0.53 1.1028 1.1028 0 0 0

0.54 1.1112 1.1112 0 0 0

0.55 1.1200 1.1200 0 0 0

0.56 1.1293 1.1293 0 0 0

0.57 1.1392 1.1392 0 0 0

0.58 1.1495 1.1495 0 0 0

0.59 1.1598 1.1598 0 0 0

0.60 1.1700 1.1700 0 0 0

0.61 1.1807 1.1807 0 0 0

0.62 1.1921 1.1921 0 0 0

0.63 1.2031 1.2031 0 0 0

0.64 1.2127 1.2127 0 0 0

0.65 1.2200 1.2200 0 0 0

0.66 1.2257 1.2257 0 0 0

0.67 1.2312 1.2312 0 0 0

0.68 1.2357 1.2357 0 0 0

0.69 1.2388 1.2388 0 0 0

0.70 1.2400 1.2400 0 0 0

0.71 1.2388 1.2388 0 0 0

0.72 1.2356 1.2356 0 0 0

0.73 1.2310 1.2310 0 0 0

0.74 1.2256 1.2256 0 0 0

0.75 1.2200 1.2200 0 0 0

0.76 1.2135 1.2135 0 0 0

0.77 1.2053 1.2053 0 0 0

0.78 1.1958 1.1958 0 0 0

0.79 1.1856 1.1856 0 0 0

0.80 1.1750 1.1750 0 0 0

0.81 1.1638 1.1638 0 0 0

Page 60: N McNamara - espace.library.uq.edu.au

N McNamara

53

0.82 1.1515 1.1515 0 0 0

0.83 1.1383 1.1383 0 0 0

0.84 1.1244 1.1244 0 0 0

0.85 1.1100 1.1100 0 0 0

0.86 1.0954 1.0954 0 0 0

0.87 1.0805 1.0805 0 0 0

0.88 1.0649 1.0649 0 0 0

0.89 1.0484 1.0484 0 0 0

0.90 1.0304 1.0304 0 0 0

0.91 1.0107 1.0000 0 0 0

0.92 0.9886 1.1938 0.1826 0.0270 -0.0958

0.93 0.9633 1.1866 0.1883 0.0308 -0.1041

0.94 0.9347 1.1789 0.1941 0.0341 -0.1093

0.95 0.9031 1.1715 0.1999 0.0369 -0.1132

0.96 0.8684 1.1650 0.2054 0.0392 -0.1173

0.97 0.8305 1.1593 0.2105 0.0407 -0.1230

0.98 0.7870 1.1540 0.2153 0.0419 -0.1290

0.99 0.7384 1.1490 0.2199 0.0428 -0.1350

1.00 0.6856 1.1443 0.2242 0.0437 -0.1410

1.01 0.6292 1.1400 0.2282 0.0445 -0.1470

1.02 0.5682 1.1360 0.2319 0.0455 -0.1531

1.03 0.4943 1.1322 0.2359 0.0467 -0.1599

1.04 0.4137 1.1286 0.2397 0.0479 -0.1669

1.05 0.3351 1.1252 0.2429 0.0490 -0.1732

1.06 0.2672 1.1220 0.2447 0.0498 -0.1782

1.07 0.2160 1.1189 0.2449 0.0502 -0.1815

1.08 0.1711 1.1159 0.2442 0.0505 -0.1844

1.09 0.1309 1.1131 0.2429 0.0507 -0.1869

1.10 0.0959 1.1104 0.2412 0.0509 -0.1888

1.11 0.0670 1.1080 0.2391 0.0510 -0.1899

1.12 0.0438 1.1057 0.2366 0.0510 -0.1899

1.13 0.0232 1.1036 0.2323 0.0509 -0.1894

1.14 0.0086 1.1020 0.2265 0.0507 -0.1885

1.15 0.0031 1.1007 0.2199 0.0505 -0.1873

1.16 0.0004 1.1000 0.2132 0.0502 -0.1858

1.17 0.0004 1.1007 0.2068 0.0498 -0.1839

1.18 0.0031 1.1020 0.2001 0.0491 -0.1802

1.19 0.0086 1.1036 0.1931 0.0481 -0.1751

1.20 0.0232 1.1057 0.1859 0.0470 -0.1692

1.21 0.0438 1.1080 0.1786 0.0457 -0.1630

1.22 0.0670 1.1104 0.1713 0.0443 -0.1569

1.23 0.0959 1.1131 0.1636 0.0424 -0.1500

1.24 0.1309 1.1159 0.1558 0.0402 -0.1425

1.25 0.1711 1.1189 0.1482 0.0380 -0.1350

Page 61: N McNamara - espace.library.uq.edu.au

N McNamara

54

1.26 0.2160 1.1220 0.1412 0.0359 -0.1281

1.27 0.2672 1.1252 0.1350 0.0342 -0.1221

1.28 0.3351 1.1286 0.1293 0.0327 -0.1167

1.29 0.4137 1.1322 0.1240 0.0313 -0.1116

1.30 0.4943 1.1360 0.1190 0.0299 -0.1068

1.31 0.5682 1.1400 0.1143 0.0286 -0.1022

1.32 0.6292 1.1443 0.1098 0.0273 -0.0978

1.33 0.6856 1.1490 0.1054 0.0259 -0.0932

1.34 0.7384 1.1540 0.1012 0.0245 -0.0888

1.35 0.7870 1.1593 0.0973 0.0232 -0.0848

1.36 0.8305 1.1650 0.0937 0.0223 -0.0814

1.37 0.8687 1.1715 0.0905 0.0218 -0.0787

1.38 0.9049 1.1789 0.0881 0.0216 -0.0760

1.39 0.9384 1.1866 0.0860 0.0215 -0.0736

1.40 0.9676 1.1938 0.0836 0.0213 -0.0717

1.41 0.9910 1.0000 0.0802 0.0209 -0.0706

1.42 1.0075 1.0075 0 0 0

1.43 1.0197 1.0197 0 0 0

1.44 1.0299 1.0299 0 0 0

1.45 1.0400 1.0400 0 0 0

1.46 1.0502 1.0502 0 0 0

1.47 1.0597 1.0597 0 0 0

1.48 1.0691 1.0691 0 0 0

1.49 1.0790 1.0790 0 0 0

1.50 1.0900 1.0900 0 0 0

1.51 1.1031 1.1031 0 0 0

1.52 1.1182 1.1182 0 0 0

1.53 1.1336 1.1336 0 0 0

1.54 1.1481 1.1481 0 0 0

1.55 1.1600 1.1600 0 0 0

1.56 1.1702 1.1702 0 0 0

1.57 1.1799 1.1799 0 0 0

1.58 1.1885 1.1885 0 0 0

1.59 1.1954 1.1954 0 0 0

1.60 1.2000 1.2000 0 0 0

1.61 1.2031 1.2031 0 0 0

1.62 1.2058 1.2058 0 0 0

1.63 1.2080 1.2080 0 0 0

1.64 1.2095 1.2095 0 0 0

1.65 1.2100 1.2100 0 0 0

1.66 1.2097 1.2097 0 0 0

1.67 1.2089 1.2089 0 0 0

1.68 1.2077 1.2077 0 0 0

1.69 1.2064 1.2064 0 0 0

Page 62: N McNamara - espace.library.uq.edu.au

N McNamara

55

1.70 1.2050 1.2050 0 0 0

1.71 1.2033 1.2033 0 0 0

1.72 1.2012 1.2012 0 0 0

1.73 1.1988 1.1988 0 0 0

1.74 1.1967 1.1967 0 0 0

1.75 1.1950 1.1950 0 0 0

1.76 1.1936 1.1936 0 0 0

1.77 1.1923 1.1923 0 0 0

1.78 1.1911 1.1911 0 0 0

1.79 1.1903 1.1903 0 0 0

1.80 1.1900 1.1900 0 0 0

1.81 1.1905 1.1905 0 0 0

1.82 1.1919 1.1919 0 0 0

1.83 1.1941 1.1941 0 0 0

1.84 1.1968 1.1968 0 0 0

1.85 1.2000 1.2000 0 0 0

1.86 1.2055 1.2055 0 0 0

1.87 1.2144 1.2144 0 0 0

1.88 1.2256 1.2256 0 0 0

1.89 1.2379 1.2379 0 0 0

1.90 1.2500 1.2500 0 0 0

1.91 1.2627 1.2627 0 0 0

1.92 1.2768 1.2768 0 0 0

1.93 1.2917 1.2917 0 0 0

1.94 1.3063 1.3063 0 0 0

1.95 1.3200 1.3200 0 0 0

1.96 1.3336 1.3336 0 0 0

1.97 1.3475 1.3475 0 0 0

1.98 1.3608 1.3608 0 0 0

1.99 1.3720 1.3720 0 0 0

2.00 1.3800 1.3800 0 0 0

2.01 1.3859 1.3859 0 0 0

2.02 1.3914 1.3914 0 0 0

2.03 1.3958 1.3958 0 0 0

2.04 1.3989 1.3989 0 0 0

2.05 1.4000 1.4000 0 0 0

2.06 1.3989 1.3989 0 0 0

2.07 1.3959 1.3959 0 0 0

2.08 1.3915 1.3915 0 0 0

2.09 1.3861 1.3861 0 0 0

2.10 1.3800 1.3800 0 0 0

2.11 1.3714 1.3714 0 0 0

2.12 1.3588 1.3588 0 0 0

2.13 1.3435 1.3435 0 0 0

Page 63: N McNamara - espace.library.uq.edu.au

N McNamara

56

2.14 1.3268 1.3268 0 0 0

2.15 1.3100 1.3100 0 0 0

2.16 1.2917 1.2917 0 0 0

2.17 1.2708 1.2708 0 0 0

2.18 1.2492 1.2492 0 0 0

2.19 1.2283 1.2283 0 0 0

2.20 1.2100 1.2100 0 0 0

2.21 1.1939 1.1939 0 0 0

2.22 1.1788 1.1788 0 0 0

2.23 1.1647 1.1647 0 0 0

2.24 1.1517 1.1517 0 0 0

2.25 1.1400 1.1400 0 0 0