Research Article Generic Model of a Satellite Attitude ...

17
Research Article Generic Model of a Satellite Attitude Control System Janusz Narkiewicz, Mateusz Sochacki , and Bartlomiej Zakrzewski Faculty of Power and Aeronautical Engineering, Warsaw University of Technology, Warsaw 00-665, Poland Correspondence should be addressed to Mateusz Sochacki; [email protected] Received 27 January 2020; Revised 30 April 2020; Accepted 20 May 2020; Published 23 July 2020 Academic Editor: Maj D. Mirmirani Copyright © 2020 Janusz Narkiewicz et al. This is an open access article distributed under the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited. A generic model of a nanosatellite attitude control and stabilization system was developed on the basis of magnetorquers and reaction wheels, which are controlled by PID controllers with selectable gains. This approach allows using the same architectures of control algorithms (and software) for several satellites and adjusting them to a particular mission by parameter variation. The approach is illustrated by controlling a satellite attitude in three modes of operation: detumbling after separation from the launcher, nominal operation when the satellite attitude is subjected to small or moderate disturbances, and momentum unloading after any reaction wheel saturation. The generic control algorithms adjusted to each mode of operation were implemented in a complete attitude control system. The control system model was embedded into a comprehensive simulation model of satellite ight. The simulation results proved the eciency of the proposed approach. 1. Introduction The advancements in miniaturization of sensors and mechanical devices and computing hardware and software lead to enhanced performance of small satellites. The novel concepts of cooperating satellites ying in a formation give prospects for a further increase of the possibility of providing services similar to those oered now by large and expensive spacecraft and also of oering new services or functionalities [1]. An important requirement for satellite operation in all ight phases, crucial for collaboration in formation ying, is eective attitude stabilization and control. Satellite control torques may be generated in a passive or active way. Passive attitude control does not consume power but may be applied only for dumping of unwanted satellite motion resulting from external disturbances [2]. Active con- trol provides higher control authority than passive methods and allows performing rotational maneuvers by using actua- tors on board a spacecraft. The most commonly used satellite actuators are magnetorquers and reaction wheels [35], sometimes also thrusters [6], but for attitude control, they are rarely used due to fuel mass and volume penalty and its irreversible depletion. When low pointing accuracy is su- cient, magnetorquers may be the only actuators used for atti- tude control [7, 8]. CubeSats attitude control systems are overviewed in [9]. Most CubeSats use reaction wheels and magnetorquers, although passive magnetic stabilization is also applied, especially in small units. After separation from a launcher, the satellite may tum- ble, i.e., spin with a high angular rate, which [1012] may be moderate, like 0.1 rad/s in each axis, but also may be above 95rad/s (600 ° /s). In nominal operating conditions, a satellite is subjected to various disturbances, which require quick and eective reactions by the attitude control system, usually using reaction wheels as actuators. Uncontrolled reaction wheels saturate, i.e., they reach the limits of maximal angular velocity, due to large external disturbances or when counter- acting disturbing torques accumulate satellite angular veloci- ties for a long time. In this research, three modes of ACS operation are considered: detumbling, nominal, and momen- tum unloading. In the detumbling mode, usually only mag- netorquers are used, as the reaction wheels may quickly saturate. However, in [4, 10, 13], reaction wheels were con- sidered also for CubeSat detumbling. In [14], B-dot and bang-bang B-dot control laws were applied for magnetor- quers, and bang-bang B-dot provided quicker stabilization Hindawi International Journal of Aerospace Engineering Volume 2020, Article ID 5352019, 17 pages https://doi.org/10.1155/2020/5352019

Transcript of Research Article Generic Model of a Satellite Attitude ...

Page 1: Research Article Generic Model of a Satellite Attitude ...

Research ArticleGeneric Model of a Satellite Attitude Control System

Janusz Narkiewicz, Mateusz Sochacki , and Bartłomiej Zakrzewski

Faculty of Power and Aeronautical Engineering, Warsaw University of Technology, Warsaw 00-665, Poland

Correspondence should be addressed to Mateusz Sochacki; [email protected]

Received 27 January 2020; Revised 30 April 2020; Accepted 20 May 2020; Published 23 July 2020

Academic Editor: Maj D. Mirmirani

Copyright © 2020 Janusz Narkiewicz et al. This is an open access article distributed under the Creative Commons AttributionLicense, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work isproperly cited.

A generic model of a nanosatellite attitude control and stabilization system was developed on the basis of magnetorquers andreaction wheels, which are controlled by PID controllers with selectable gains. This approach allows using the samearchitectures of control algorithms (and software) for several satellites and adjusting them to a particular mission by parametervariation. The approach is illustrated by controlling a satellite attitude in three modes of operation: detumbling after separationfrom the launcher, nominal operation when the satellite attitude is subjected to small or moderate disturbances, and momentumunloading after any reaction wheel saturation. The generic control algorithms adjusted to each mode of operation wereimplemented in a complete attitude control system. The control system model was embedded into a comprehensive simulationmodel of satellite flight. The simulation results proved the efficiency of the proposed approach.

1. Introduction

The advancements in miniaturization of sensors andmechanical devices and computing hardware and softwarelead to enhanced performance of small satellites. The novelconcepts of cooperating satellites flying in a formation giveprospects for a further increase of the possibility of providingservices similar to those offered now by large and expensivespacecraft and also of offering new services or functionalities[1]. An important requirement for satellite operation in allflight phases, crucial for collaboration in formation flying,is effective attitude stabilization and control.

Satellite control torques may be generated in a passive oractive way. Passive attitude control does not consume powerbut may be applied only for dumping of unwanted satellitemotion resulting from external disturbances [2]. Active con-trol provides higher control authority than passive methodsand allows performing rotational maneuvers by using actua-tors on board a spacecraft. The most commonly used satelliteactuators are magnetorquers and reaction wheels [3–5],sometimes also thrusters [6], but for attitude control, theyare rarely used due to fuel mass and volume penalty and itsirreversible depletion. When low pointing accuracy is suffi-

cient, magnetorquers may be the only actuators used for atti-tude control [7, 8]. CubeSats attitude control systems areoverviewed in [9]. Most CubeSats use reaction wheels andmagnetorquers, although passive magnetic stabilization isalso applied, especially in small units.

After separation from a launcher, the satellite may tum-ble, i.e., spin with a high angular rate, which [10–12] maybe moderate, like 0.1 rad/s in each axis, but also may be above95rad/s (600°/s). In nominal operating conditions, a satelliteis subjected to various disturbances, which require quickand effective reactions by the attitude control system, usuallyusing reaction wheels as actuators. Uncontrolled reactionwheels saturate, i.e., they reach the limits of maximal angularvelocity, due to large external disturbances or when counter-acting disturbing torques accumulate satellite angular veloci-ties for a long time. In this research, three modes of ACSoperation are considered: detumbling, nominal, and momen-tum unloading. In the detumbling mode, usually only mag-netorquers are used, as the reaction wheels may quicklysaturate. However, in [4, 10, 13], reaction wheels were con-sidered also for CubeSat detumbling. In [14], B-dot andbang-bang B-dot control laws were applied for magnetor-quers, and bang-bang B-dot provided quicker stabilization

HindawiInternational Journal of Aerospace EngineeringVolume 2020, Article ID 5352019, 17 pageshttps://doi.org/10.1155/2020/5352019

Page 2: Research Article Generic Model of a Satellite Attitude ...

but with a higher final angular velocity error than the B-dotone. In [15], satellite angular velocity feedback was imple-mented for the detumbling mode.

The nominal control laws provide stabilization counter-acting external disturbances, which, even when small (likegravity-gradient) but acting for a long time, may drift a satel-lite away from its required attitude. Another control task inthe nominal mode is satellite rotational maneuvers. As pri-mary actuators, reaction wheels are commonly used in thenominal mode, as they may assure a high attitude accu-racy and force a relatively quick response (although slowerthan thrusters). To control reaction wheels, proportional-derivative (PD) controllers are used [16–24], sometimes withmodifications like including additional nonlinear terms[16, 23]. The control laws for the nominal mode are oftendeveloped using a simplified satellite model (for instance,only rotational motion is considered neglecting disturbancetorques) before being applied to nonlinear systems [16].

In such a case, reaction wheels require diminishing theirmomentum. For small satellites in low Earth orbits (LEO),magnetorquers are used to unloadmomentum, as cheap, reli-able, and effective torque-generating devices [3]. In [24, 25],the crossproduct control law was applied to control threereaction wheels, and in [3], its modification called arevisited-product control law was investigated. A more gen-eral approach was presented in [5] to control a set of threeor more reaction wheels.

For each mode of operation, the signals from the controllaw block should be passed as input to particular actuators. Itmay be achieved in various ways, for instance, using one typeof actuators only (reaction wheels in the nominal mode ormagnetorquers in the detumbling mode) and then minimiz-ing the control effort [3, 4].

The main objective of the study was to develop an atti-tude control system to be implemented on board CubeSatsfor prospective formation flying. The ease of applicability tosimilar satellite sizes was the main requirement to diminishthe cost of control system development and tuning. Thisexplains why the traditional PID approach was implementedwith focus on flexibility of implementation to similar satel-lites and scalability for a number of actuators. To evaluatefeasibility of the developed concept, a comprehensive WUTsimulation model of a satellite flight was supplemented by acontrol system model, allowing evaluation of the controlquality. The paper will describe the satellite simulationmodel, and then next, the generic structure of the control sys-tem will be presented. The simulation results will confirm theefficiency of the proposed control system and algorithms.

2. Satellite Simulation Model

The satellite simulation software developed at the WarsawUniversity of Technology (Figure 1) contains satellitedynamic and kinematic equations, models of sensors andactuators, and algorithms of the navigation and control sys-tem. The modular structure allows flexibility in model exten-sion and modifications. The model operates in the timedomain.

A satellite bus is modeled as a rigid body with sixdegrees of freedom and time-varying mass and inertia.The dynamic equations of motion are derived in the BodyFrame Fixed (BFF) coordinate system (Figure 2), the kine-matic equations are derived in the Earth Centered Inertial(ECI) coordinate system, and the control loads are calcu-lated first in local actuator coordinate systems and thentransferred to BFF (Figure 2).

The rigid spacecraft dynamic equations of motion aredescribed as

A tð Þ _x + B t, xð Þx|fflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflffl}Inertia loads

= fG + fA + fSRP + fM|fflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflffl}External loads

+ fR + fT + fQ|fflfflfflfflfflfflffl{zfflfflfflfflfflfflffl}Control loads

, ð1Þ

where t is time, x = ½U V W P Q R�T is the satellite statevector composed of components of satellite linear vF and

angular ωF velocity vectors, and fð Þ = FTð Þ MTð Þ

h iTis the

load vectors composed of force FTð Þ and momentMTð Þ compo-

nents. Indices refer to the following: G: gravity; A: aerody-namics; SRP: solar radiation pressure; M: magnetic field(acting on the bus); R: reaction wheels (or other momentumexchange devices); T : propulsion (not used in this study); Q:magnetorquers.

The inertia matrix AðtÞ has the form

A tð Þ =

m 0 0 0 Sz −Sy0 m 0 −Sz 0 Sx

0 0 m Sy −Sx 0

0 −Sz Sy Ixx −Ixy −IxzSz 0 −Sx −Ixy Iyy −Iyz−Sy Sx 0 −Ixz −Iyz Izz

2666666666664

3777777777775, ð2Þ

wherem, Sx, Sy, Sz , Ixx, Iyy , Izz , Ixy, Iyz , and Ixz are the satellitemass, static mass moments, and moments and products ofinertia, respectively.

The Bðt, xÞ matrix is calculated as

B t, xð Þ =O xð ÞA tð Þ, ð3Þ

where

O xð Þ =

0 −R Q 0 0 0

R 0 −P 0 0 0

−Q P 0 0 0 0

0 −W V 0 −R Q

W 0 −U R 0 −P

−V U 0 −Q P 0

2666666666664

3777777777775: ð4Þ

2 International Journal of Aerospace Engineering

Page 3: Research Article Generic Model of a Satellite Attitude ...

The satellite kinematic equations for position and atti-tude have the form

_rI = CIFvF ,

_qI =12QqI ,

ð5Þ

where rFI = xFI yFI zFI½ �T is the satellite position in theinertial coordinate system (ECI), vF = U V W½ �T is thesatellite linear velocity in the satellite body frame, ωF =P Q R½ �T is the satellite angular velocity in the satellitebody frame, qI = q0 q1 q2 q3½ �T is the quaternion of asatellite attitude in the inertial frame, and CI

F is the matrixof transformation from the satellite body to inertial framein the form

CIF = CI

F qð Þ =q20 + q21 − q22 − q23 2 q1q2 − q3q0ð Þ 2 q1q3 + q2q0ð Þ2 q1q2 + q3q0ð Þ q20 − q21 + q22 − q23 2 q2q3 − q1q0ð Þ2 q1q3 − q2q0ð Þ 2 q2q3 + q1q0ð Þ q20 − q21 − q22 + q23

2664

3775:

ð6Þ

Matrix Q is composed of satellite angular velocities andhas the form

Q =

0 −P −Q −R

P 0 R −Q

Q −R 0 P

R Q −P 0

2666664

3777775: ð7Þ

It is assumed that on a satellite, N control actuators aremounted, producing torques MCi only, so the total controltorque is

fC = 〠N

i=1fCi− = 0 〠

N

i=1MCi

" #T

  i = 1,⋯,Nð Þ, ð8Þ

where MCi is the torque generated by the i-th actuator(reaction wheel or magnetorquer) in the BFF coordinatesystem.

A reaction wheel is modelled as a flywheel fixed to thesatellite body spinning with angular velocityΩR and moment

Conrol loads:(i) Reaction

wheels. (ii) Magnetorquers

(iii) Thrusters

Satellite dynamics:Loads:(i) Inertia loads(ii) Gravity(iii) Aerodynamics(iv) Solar radiation(v) Magnetic loads

Controlalgorithm

Guidancealgorithm

Sensors:(i) Magnetometers(ii) Gyroscopes(iii) Accelerometers(iv) Sun sensors(v) Earth sensors(vi) Star trackers(vii) GPS receiver(viii) Others

Navigationalgorithm

Satellitekinematics

Figure 1: Satellite simulation model architecture.

xI

xB

xF

zF

yF

yI

yB

zI

OB

OI

rFI

zB

OF

rCM

Figure 2: Coordinate system applied.

3International Journal of Aerospace Engineering

Page 4: Research Article Generic Model of a Satellite Attitude ...

of inertia IRx . Its angular momentum in the actuator coordi-nate system is

HR = IRx ΩR + ωRxð Þ 0 0½ �T , ð9Þ

where HR is the angular momentum of the reaction wheel,IRx is the moment of inertia of the reaction wheel about itsspin axis, ΩR is the spin velocity of the reaction wheel, andωR = ωRx ωRy ωRz

� �T is the satellite angular velocity inthe actuator coordinate system.

A torque acting on the satellite due to a single reactionwheel (in the BFF frame) is calculated as

MR = −IRx _ΩR + _ωRx

� �−IRxωRz ΩR + ωRxð Þ IRxωRy ΩR + ωRxð Þ

h iT:

ð10Þ

Angular velocity ΩR of a reaction wheel is the controlvariable for the generated torque.

The reaction wheel is driven by an electric motor, thedynamics of which in general is described by combined elec-tromechanical equations. As control signals, COTS (com-mercial off-the-shelf) reaction wheels use current, torque,or angular velocity, and their control electronics convertsthem to the input voltage of the motor. In this study, it isassumed that the armature current ia is a control variable,and the reaction wheel model is described by mechanicalstate equations in the form

IRx _ΩR + _ωRx

� �= kRiR − bRΩR, ð11Þ

where iR is the armature current, kR is the motor torque con-stant, and bR is the viscous friction coefficient of the motorshaft.

The torque from a magnetorquer is a crossproduct of themagnetic dipole generated by a coil and the Earth magneticfield vector:

MQ =DQ × BQ = 0 −DQBQz DQBQy� �T , ð12Þ

where DQ = DQ 0 0� �T is the magnetic dipole generated

by the coil and BQ = BQx BQy BQz� �T is a vector of the

Earth magnetic field in the magnetorquer coordinate system.A magnetic dipole is calculated as

DQ =D0uQ, ð13Þ

where D0 is the nominal dipole moment of the magnetorquerat a given nominal voltage and uQ is the control variable of amagnetorquer.

3. Attitude Control System

The general architecture of a satellite attitude control systeminvestigated in this study is shown in Figure 3. The blocks“actuators” and “satellite” (Figure 3) are the models describedin the previous chapter used for simulation of bus and actu-

ators dynamics. “Sensors” are not considered in this study.The block “current state” data contains state variables of asatellite, reaction wheels and magnetorquers, quaternionqI = q0 q1 q2 q3½ �T representing satellite attitude inthe inertial frame, vector ωF = P Q R½ �T of a satelliteangular velocity expressed in the satellite body coordinatesystem, vector ΩRW of reaction wheels angular velocities,and vector uM of the magnetorquer control signal. Therequired values of these quantities, denoted with index cmd,are taken from the block named “flight management.”

The differences between the required and current states areinputs to the satellite attitude control system (ACS), composedof two blocks denominated in Figure 3 as “satellite attitudecontrol” and “reaction wheel desaturation.” The task of the sat-ellite attitude control block is to bring a satellite to the requiredattitude state and/or to keep it at that state. The task of the reac-tion wheel desaturation block is to decelerate them to angularvelocities determined by vector Ωcmd. The desaturation pro-cess should not influence significantly the satellite rotationaldynamics, so it is included as a separate line of control.

For both control blocks, full PID control is envisaged andimplemented in the algorithm.

u ∗ð Þ =KP ∗ð Þσ ∗ð Þ +KD ∗ð Þ _σ ∗ð Þ +KI ∗ð Þ

ðσ ∗ð Þdt, ð14Þ

where σ is the vector of the differences between the requiredand current values; KPð∗Þ, KDð∗Þ, and KIð∗Þ are the propor-tional, derivative, and integral gain matrices, respectively;and ð∗Þ is the index of the actuator used.

The applied PID control is illustrated in Figure 4 showingthe selection of actuators and control gain matrix for threesystem operating modes: detumbling, nominal, and momen-tum unloading.

3.1. Detumbling Mode. The detumbling mode is activatedafter separation of the satellite from the launch vehicle or atany time when the satellite’s angular velocity is too large.The satellite angular velocity is reduced by the attitudecontroller, and the momentum management controller isinactive. In the detumbling mode, only magnetorquersare used for spacecraft stabilization, and PID is reducedto differential control related to the actual value of satelliteangular velocity

uQ =KDQωe =KDQω, ð15Þ

as the required satellite velocity reduction ωcmd = 0.The gain matrix KDQ depends on the magnetorquer

arrangement, magnetic field vector, satellite inertia, andorbital period (similar to [15]) and is calculated as

KDQ =4πTorb

I� �D−1

0 G+MTQB×

F

BFj j2 , ð16Þ

where NQ is the number of magnetorquers, D0 = diagD01 ⋯ D0NQ

is the diagonal matrix composed of

4 International Journal of Aerospace Engineering

Page 5: Research Article Generic Model of a Satellite Attitude ...

nominal dipole moments D0i of magnetorquers, GMTQ isthe matrix (3 ×NQ) representing the configuration ofmagnetorquers in the satellite body frame (the “+” in thesuperscript indicates a pseudoinverse or an inverse ofmatrix GMTQ if it exists), BF is the magnetic field vector(3 × 1) in BFF (the “×” in the superscript indicates a skewsymmetric matrix (3 × 3) satisfying B×

Fω = BF × ω), Torb isthe orbital period of the satellite, and I is the matrix ofthe satellite moment of inertia matrix about its centerof mass.

When the specified low angular velocity is achieved, theattitude control system is switched from the detumblingmode to the more precise nominal mode.

3.2. Nominal Mode. The nominal mode is used to maintainor change satellite attitude as required with adequate accu-racy, despite external disturbances. The control is performedby reaction wheels in a full PID form

uR =KPRσR +KDRωe +KIR

ðσRdt: ð17Þ

In (17), the attitude error vector σ is calculated as a rota-tion from the actual attitude qI = q0 q1 q2 q3½ �T to thecommanded attitude qIcmd = q0c q1c q2c q3c½ �T in a qua-ternion notation as [24]

q0e

q1e

q2e

q3e

2666664

3777775 =

q0c q1c q2c q3c

−q1c q0c q3c −q2c−q2c −q3c q0c q1c

−q3c q2c −q1c q0c

2666664

3777775

q0

q1

q2

q3

2666664

3777775,

σR = 2q0e q1e q2e q3e½ �T :

ð18Þ

The derivative of attitude error _σR is the difference of sat-ellite angular velocities:

ωFe = ωFcmd − ωF : ð19Þ

Flight managementCurrentrequired

state

Attitude control system Satellite model

+

+

+

–Satellite attitude

controlActuators Satellite

SensorsCurrent state

Reaction wheelsdesaturation

Figure 3: General architecture of the attitude control system.

Detumbling mode

Attitude control

MTQ P I D

RW P I D

MTQ P I D

RW P I D

MTQ P I D

RW P I D

MTQ P I D

RW P I D

Nominal mode

Attitude control

Momentumunloading mode

Attitude control

Reaction wheeldesaturation

Figure 4: Attitude control system modes of operation.

5International Journal of Aerospace Engineering

Page 6: Research Article Generic Model of a Satellite Attitude ...

The gain matrices KPR, KDR, and KIR are calculated as

KPR = ω2n +

2ζωn

T

� �R−1G+

RWI, ð20Þ

KDR = 2ζωn +1T

� �R−1G+

RWI, ð21Þ

KIR =ω2n

TR−1G+

RWI, ð22Þ

where NR is the number of reaction wheels; R = diagkR1 ⋯ kRNR

is the diagonal matrix (NR ×NR) contain-

ing reaction wheel motor constants; GRW is the matrix(3 ×NR) representing the configuration of reaction wheels,relative to satellite body axes; and ωn, ζ, and T are therequired linear control bandwidth, dumping ratio, and timeconstant of the integral controller, respectively.

3.3. Momentum Unloading Mode. The momentum unload-ing mode is used to prevent reaction wheel saturation byreducing their angular velocities. In the momentum unload-ing mode, reaction wheels are used for satellite attitude con-trol, and simultaneously, magnetorquers are switched on forreaction wheel desaturation. The mode is activated when thevelocity of at least one reaction wheel exceeds its design limit.However, for better mission performance, the momentumunloading system may be switched on from time to time toreduce accumulating wheel rotating velocity.

In the momentum unloading mode, the control of reac-tion wheels should assure satellite attitude control anddiminish the spin of reaction wheels, so the total reactionwheel control uR in this mode is done as

uR = uRS + uRR, ð23Þ

where uRS is applied to perform the first task and uRR the sec-ond one.

As the integral term in PID might slow down the controlreaction, for satellite attitude control uRS in the momentumunloading mode, the PD controller for reaction wheels isused:

uRS =KPRSσR +KDRSωe, ð24Þ

with modified P and D gain matrices (eq. (20) and (21)) cal-culated here as

KPRS = ω2nR−1G+

RWI, ð25Þ

KDRS = 2ζωnR−1G+RWI: ð26Þ

To decelerate angular velocities of reaction wheels to aspecified value Ωcmd, uRR is calculated as

uRR =KDRRΩe, ð27Þ

where Ωe =Ωcmd −ΩRW is the reaction wheel angular veloc-ity error vector, and matrix KRR is calculated as

KRR = kRR−1IRW , ð28Þ

where kR is the scalar control gain.Themagnetorquers are also controlled by the D controller:

uQ =KDQΩe, ð29Þ

and gain KDQ is calculated as [3, 24, 25]

KDQ =kQBFj j2 D

−10 G+

MTQB×FGMTQIRW , ð30Þ

where kQ is the scalar control gain.Proper selection of control gains kQ and kR in the

momentum unloading mode results in lower disturbancesof spacecraft attitude and shorter duration of the reactionwheel desaturation process.

4. Simulation Study

To investigate the efficiency of the developed attitude controlsystem, a simulation study was done for the generic satellitemass data given in Table 1.

The initial satellite orbit elements are provided in Table 2.

Table 1: Satellite mass properties.

Description Value Unit

Mass 6.2 [kg]

Moments of inertia: Ixx , Iyy , Izz 0.0756, 0.0763, 0.0209 [kg·m2]

Products of inertia: Ixy , Iyz , Ixz -0.0002, 0.0020, -0.0019 [kg·m2]

Location of center of mass relative to geometrical center (x, y, z) 0.0034, -0.0036, -0.0007 [m]

Dimensions 0:7 × 0:7 × 0:4 [m]

Table 2: Initial satellite orbital parameters.

Description Value Unit

Semimajor axis 6371 000 [m]

Eccentricity 0 [-]

Longitude of ascending node 331.7739 [°]

Orbit inclination 51.6393 [°]

Argument of perigee 0 [°]

True anomaly 20.5285 [°]

Orbital period 5500 s

6 International Journal of Aerospace Engineering

Page 7: Research Article Generic Model of a Satellite Attitude ...

Three configurations of reaction wheels: orthogonal,pyramidal, and tetrahedral (Figure 5), were investigatedaccompanied by a set of three orthogonal magnetorquers.

The reaction wheel parameters used in simulations arebased on Sinclair RW-0.03 data [26] and are given in Table 3.

The major parts of simulations were performed for theorthogonal configuration of reaction wheels as the mostcommonly used one [9]. The other two configurations wereinvestigated mainly for validating the scalability of the devel-oped algorithms.

The matrices GRW describing reaction wheel configura-tion, used in equations (20), (21), (22) and (25), (26), (27) are

Orthogonal configuration: identity matrix dimension(3 × 3)

Pyramid configuration:

GRW =

0:5774 −0:5774 −0:5774 0:5774

0:5774 0:5774 −0:5774 −0:5774

0:5774 0:5774 0:5774 0:5774

2664

3775 ð31Þ

Tetrahedron configuration:

GRW =

0:9428 −0:4714 −0:4714 0

0 0:8165 −0:8165 0

−0:3333 −0:3333 −0:3333 1

2664

3775: ð32Þ

The set of magnetorquers contains three devices in anorthogonal configuration with nominal dipoles as in [27]:in the OFxF and OFyF axes, D01 =D02 = 0:2Am2, and inthe OFzF axis, D03 = 0:24Am2.

For the detumbling mode, the constant part of the gainmatrix was calculated purely based on the satellite inertiaand orbital parameters as per equation (16). The selectiveparameters for the nominal (equations (20), (21), (22)) andmomentum unloading (equations (28) and (30)) modeswere chosen using the trial and error method. The attitudecontroller parameters were set to ωn = 0:8 rad/s, ζ = 1, andT = 12:5 s. For the nominal mode, the aim was to allowrelatively quick maneuvers without significant overshootof the attitude and sufficiently high pointing accuracy forthe initial wheel rates of 0 rpm/s. The gains in the momen-tum unloading mode, kQ = kR = 0:05, were chosen to have

relatively small values in order to limit the deviation fromthe required attitude during that phase.

Initial attitudes for all simulated cases of yaw, pitch, androll angle values were zero, i.e., Ψ0 =Θ0 =Φ0 = 0.

The simulation rationale for all cases was to check theapplicability of the algorithms developed. In particular, forthe detumbling mode, the ability to suppress the satelliterotation was verified. For the normal operating mode, theability to maintain or change the satellite attitude was inves-tigated, for various angular velocities and configurations ofreaction wheels. For the momentum unloading case, the abil-ity to desaturate high rates of reaction wheels was verified.

4.1. Detumbling Mode. The detumbling mode of the ACSoperation was simulated to verify the effectiveness of reduc-ing the satellite’s high value of 10°/s per axis of initial angularvelocity, which may appear after separation from thelauncher. The magnetorquers were actuators used in thismode. The time profiles of the satellite angular velocity mod-ule and magnetorquer dipole moments module are presentedin Figures 6(a) and 6(b).

The satellite angular velocity was reduced below 0.5°/safter 3200 s and below 0.2°/s after 6000 s, i.e., in time about9% longer than the orbital period which was 5500 s. Theresult is satisfying, in terms, for instance, of the require-ment formulated in [10] that a satellite should detumblein about one orbital period from the initial angular veloc-ity of 5.73°/s per axis.

As indicated by other simulations, using the detumblingmode controller did not reduce the satellite rate below0.2°/s. To cross this threshold, the ACS nominal operationmode should be used.

4.2. Nominal Mode. In the simulations of the nominal mode,only reaction wheels were active. The quantities monitoredfor this mode are satellite Euler angles: Ψ (yaw), Θ (pitch),

x

z

y

(a)

x y

z

(b)

x

z

y

(c)

x

y

z

(d)

Figure 5: Reaction wheel configurations: (a) orthogonal, (b) pyramidal, (c) tetrahedral, and (d) skew.

Table 3: RW data.

Description Value Unit

Moment of inertia about spin axis 51.16 [kg·mm2]

Armature current 0.35 [A]

Viscous friction 3.837 [kg·mm2/s]

Motor torque constant 51.6393 [°]

7International Journal of Aerospace Engineering

Page 8: Research Article Generic Model of a Satellite Attitude ...

and Φ (roll), angular rates of reaction wheels, and optionallysatellite angular velocity. The list of scenarios simulated forthe nominal mode of ACS operation is given in Table 4.

The simulation of case 1 illustrates the transition fromdetumbling to nominal mode. The initial angular velocityvector is assumed to be 0.5°/s per axis, which is the residualof the detumbling mode. The requirement was to keep thezero Euler angles, i.e., Ψ = 0°, Θ = 0°, and Φ = 0°.

At the beginning of the control, there was an increase inthe reaction wheel angular rate, but after 10 seconds(Figure 7), the satellite angular rate was reduced from 0.5°/sto almost 0°/s in each axis, which was the result of reactionwheel activity. At that moment, the velocities of the firstand the second reaction wheels were stabilized at 120 rpm.The third wheel stabilized at 30 rpm, because the satellitemoment of inertia about the respective axis was lower thanthose on X and Y axes which are approximately equal.

The simulation case 2 was performed to investigate atti-tude control system performance for different initial reactionwheel rates. In this case, the initial attitude wasΨ = 0°,Θ = 0°,andΦ = 0° and initial angular velocity vector was 0°/s per axis.The satellite was required to keep this status for 100 s and thenchange attitude toΨ = 120°,Θ = 60°, andΦ = 150°. The simu-

lation was performed for a set of 3 orthogonal reaction wheels,for 3 different initial velocities of reaction wheels.

In case 2 for initial RW rates 0 and 2000 rpm, the perfor-mance is acceptable. The high initial reaction wheel rate suchas 4000 rpm does not allow achieving a large commandedangle of rotation, which may be the result of reaction wheelsaturation, as the second reaction wheel reached the maximalangular velocity of 6200 rpm (Figure 8). The remedy is thatfor the required large satellite attitude variation, the maneu-ver should be divided into a few smaller attitude changes toavoid high angular velocity of reaction wheels. Another solu-tion could be to use different values of bandwidth and dump-ing in the gain matrices for this controller.

The case 3 simulation was performed to check the ACSperformance for the cases similar for a time-varying nadircommanded attitude. In the first 100 seconds, the com-manded attitude was Ψ = 0°, Θ = 0°, and Φ = 0°. After that,the commanded satellite attitude was varying for nadir point-ing. The simulation was performed for 3 different sets of ini-tial reaction wheel velocities (0, 2000, and 4000 rpm). Theresults are presented in Figures 9(a) and 9(b).

In this case, the control system response is similar to thatin case 2. The system has a better performance for lower

0

20

15

10

5

0

0.4

0.3

0.2

0.1

0

1000

|ω| (

º/s)

|D| (

Am

2 )

2000 3000 4000 5000 6000

0 1000 2000 3000

t (s)

4000 5000 6000

(a)

0.2

0.1

04000

×10−3

4

3

2

1

04000 4500 5000

t (s)5500 6000

4500 5000 5500 6000

|ω|(º

/s)

|D| (

Am

2 )

(b)

Figure 6: (a) Satellite rate reduction in the detumbling mode; (b) closeup.

Table 4: Nominal mode simulation cases.

Case Initial satellite rates P =Q = RCommanded attitude

(Initial)(After 100 s)

Initial RW spin RW configuration

1 0.5°/s0°

0°0 Orthogonal 3 wheels

2 00°, 0°, 0°

120°, 60°, 1500, 2000, 4000 rpm Orthogonal 3 wheels

3 00°, 0°, 0°

Nadir pointing0 rpm Orthogonal 3 wheels

4 0 120°, 60°, 150° 0 rpmPyramid 4 wheels

Tetrahedron 4 wheels

8 International Journal of Aerospace Engineering

Page 9: Research Article Generic Model of a Satellite Attitude ...

0

0.2

0.1

0

12080

𝛺i (

rpm

)𝛷, 𝛩

, 𝛹 (º)

P, Q

, R (º/s)

40

0.4

0.2

0

2 4 6 8 10 12 14 16 18 20

0 2 4 6 8 10 12 14 16 18 20

t (s)

00 2 4 6 8 10 12 14 16 18 20

P

RW1RW2RW3

Q

R

𝛷

𝛩

𝛹

(a)

10–1

0

1

1

0

−1

20

×10−3

×10−4

30 40 50 60 70 80 90 100

10

120

80

40

0

20 30 40 50 60 70 80 90 100

10 20 30 40 50

t (s)60 70 80 90 100

𝛺i (rp

m)

𝛷, 𝛩

, 𝛹 (º)

P, Q

, R (º/s)

P

RW1

RW2

RW3

Q

R

𝛷

𝛩

𝛹

(b)

Figure 7: (a) Stabilization in the nominal mode, case 1; (b) closeup.

9International Journal of Aerospace Engineering

Page 10: Research Article Generic Model of a Satellite Attitude ...

initial reaction wheel rates, as the commanded attitude wasreached more quickly and with better accuracy. For initialrates of 4000 rpm, the commanded attitude is not achievedand a significant pointing error occurs. It may be a symptomof reaction wheel saturation, as a result of summing up theangular velocities of reaction wheels and satellite, whichmakes the controller ineffective.

The simulation case 4 was performed to verify whetherthe developed control laws may be also used for reactionwheel configurations other than the orthogonal one. Initialsatellite angular velocity vector was 0°/s per axis. The pyra-mid and tetrahedron configurations of four reaction wheelswere investigated. In control gain calculations, the configura-tions differ by matrices GRW . In these cases, the satellite wasexpected to change attitude from initial Ψ = 0°, Θ = 0°, andΦ = 0° to the commanded one: Ψ = 120°, Θ = 60°, and Φ =150°. The results are presented in Figures 10(a) and 10(b).

For both configurations, the maneuver was completedafter about 15 seconds. The satellite motion was differentfor each configuration, but the final results were similar.The overshoot for pyramid configuration was larger on thepitch axis than for the tetrahedron configuration, but onthe roll axis, an opposite tendency occurred.

The systems using 4 reaction wheels reach the com-manded attitude significantly earlier with smaller overshootthan the one with 3 reaction wheels (case no. 3), as usingmore reaction wheels allows to generate a higher torque onthe satellite. The performance of reaction wheel arrays is dif-ferent; however, in both cases, the maximal speed achievedby any reaction wheel was similar (above 2000 rpm for the1st wheel of the pyramid configuration and for the 2nd wheelof the tetrahedron configuration).

The results presented in Figure 10(a) illustrate that fordifferent configurations, the performance of the control sys-temmay be similar. This poses a difficult task for the designerwhich configuration should be chosen if 4 reaction wheels areused. A wider set of scenarios would indicate more differ-

ences between those 2 configurations. However, a detailedcomparison between different reaction wheel configurationsis not covered in this study.

4.3. Momentum Unloading Mode. In this mode of ACS oper-ation, both reaction wheels and magnetorquers are active. Toverify the effectiveness of diminishing reaction wheel angularvelocities, a desaturation process was performed for initialvelocities of reaction wheels equal to 6000 rpm which is neartheir maximal value of operation—6200 rpm. Although it is arather unrealistic value in practice, this case was studied tocheck the efficiency of the control. Simulations were donefor orthogonal and pyramid configurations of reactionwheels.

For the orthogonal configuration of three reaction wheels(Figures 11(a) and 11(b)), their spin velocities were reducedto near zero values after 6400 s, i.e., 115% of the orbitalperiod, which is a satisfying result. During the control, therates of reaction wheels were not decreasing monotonically.It may be attributed to the fact that magnetorquers wereoperating close to their saturation region for most of thetime. The satellite pointing error was decreasing withdecreasing rates of reaction wheels. It may be explained byviscous dumping of reaction wheel rotation (providing resid-ual torque) which was not compensated for in the momen-tum unloading mode.

To verify whether it is possible to apply control lawsdeveloped during this study to more than three reactionwheels, the simulation for a set of 4 reaction wheels in thepyramid configuration was performed.

The desaturation process was successful for this case too(Figures 12(a) and 12(b)), as angular rates of all four reactionwheels were reduced to near zero. It took more time com-pared to the operation of 3 orthogonal reaction wheels, mostlikely due to the fact that a system of 4 reaction wheels has ahigher total angular momentum to be reduced. This mightalso result in a higher initial pointing error.

160

𝛷 (°

)𝛩

(°)

𝛺2

(rpm

)𝛺

3 (r

pm)

𝛺1

(rpm

)

𝛹 (°

)

12080

–80

40

–40

40

0

0

0

0

00100

–100

4000

6000

2000

–2000

–2000

4000

4000

2000

2000

–2000–4000

100 100150 150200 200250 250300 300t (s)t (s)

cmd0 rpm

2000 rpm4000 rpm

Figure 8: Nominal mode, case 2.

10 International Journal of Aerospace Engineering

Page 11: Research Article Generic Model of a Satellite Attitude ...

4.4. Simulation Result Summary. The objective of the simula-tion study was to verify whether the attitude control systemdeveloped on the basis of linear control may be applied tothe nonlinear satellite model. Three modes of operationswere simulated, and the results of the control are satisfactory.

In the detumbling mode, the satellite angular velocity wasreduced even for very high initial angular velocities, such as10°/s per axis. The other simulation results indicated thatusing only magnetorquers for high accuracy pointing in thepresence of external torque acting on the satellite may notbe effective, which suggests that reaction wheels should beused in the final tuning of satellite attitude.

The simulations for the nominal mode of operation indi-cated that the control algorithm, applied in this mode, pro-vides sufficient attitude control. The control algorithmdeveloped in this study may be applied to various configura-tions of reaction wheels, which was illustrated by simulationresults for orthogonal, pyramid, and tetrahedron reactionwheel configurations. The study revealed limits of using reac-tion wheels at high angular rates. As varying external loadsacting on a satellite results in increase of reaction wheel angu-lar velocities, the momentum unloading mode is necessary tomaintain the required performance of reaction wheel actua-tion during a mission.

0–40–80

–120

0–20–40–60–80

120

80

40

0 –6000–4000–2000

020004000

0

2000

4000

60000

2000

4000

6000

100 120 140t (s) t (s)

160 180 200 100 120 140 160 180 200

𝛺3 (

rpm

)𝛺

2 (rp

m)

𝛺1 (

rpm

)

𝛹 (º

)𝛩

(º)

𝛷 (º

)

cmd0 rpm

2000 rpm4000 rpm

(a)

140

102

110

118

–49

–45

–41

–129

–125

–121

150

cmd0 rpm

2000 rpm4000 rpm

160 170t (s)

180 190 200

𝛹 (º

)𝛩

(º)

𝛷 (º

)

(b)

Figure 9: (a) Nominal mode, case 3; (b) closeup.

11International Journal of Aerospace Engineering

Page 12: Research Article Generic Model of a Satellite Attitude ...

120

80

𝛷 (º

)

𝛺1 (r

pm)

𝛺2 (r

pm)

𝛺3 (r

pm)

𝛺4 (r

pm)

𝛩 (º

)𝛹

(º)

40

0

60

40

20

0120

80

40

0 0

500

1000

0

1000

2000

–2000

–1000

–2000

–1000

0

0

0 10 20t (s)

30 40 50

cmdPyramidTetrahedron

0 10 20t (s)

30 40 50

(a)

150.4

149.6

60.4

59.6

120.4

120

119.6

15 20 25 30 35 40 45 50

cmdPyramidTetrahedron

60

150

t (s)15 20 25 30 35 40 45 50

t (s)

0

–10

–20

0

–10

–20

0

–10

–20

0–10

–20

𝛹 (º

)𝛩

(º)

𝛷 (º

)

𝛺1 (

rpm

)𝛺

2 (rp

m)

𝛺3 (

rpm

)𝛺

4 (rp

m)

(b)

Figure 10: (a) Nominal mode, case 4; (b) closeup.

12 International Journal of Aerospace Engineering

Page 13: Research Article Generic Model of a Satellite Attitude ...

6000

4000

2000

0

0.4

0.3

0.2

0.1

025201510

50

0 2000 4000 6000 8000 10000

𝛷, 𝛩

, 𝛹 (º

)|D

| (A

m2 )

𝛺i (

rpm

)

t (s)

RW (1)RW (2)RW (3)

𝛷𝛹

𝛩

(a)

RW (1)RW (2)RW (3)

𝛷𝛹

𝛩

0.4

240

–240

0

0.3

0.2

0.1

00.08

–0.08

0

6000 7000 8000 9000 10000

𝛷, 𝛩

, 𝛹 (º

)|D

| (A

m2 )

𝛺i (

rpm

)

t (s)

(b)

Figure 11: (a) Momentum unloading mode, reaction wheels in orthogonal configuration. (b) Closeup.

13International Journal of Aerospace Engineering

Page 14: Research Article Generic Model of a Satellite Attitude ...

6000

4000

2000

0

0.4

0.3

0.2

0.1

0252015

𝛷, 𝛩

, 𝛹 (º

)|D

| (A

m2 )

𝛺i (

rpm

)

105

00 2000 4000 6000

t (s)8000 10000 12000

RW (1)RW (2)

RW (3)RW (4)

𝛷

𝛹

𝛩

(a)

240

0

–240

0.4

0.3

0.2

0.1

0

0.04

–0.04

9000 10000 11000 12000

0

𝛺i (

rpm

)𝛷

, 𝛩, 𝛹

(º)

|D| (

Am

2 )

t (s)

RW (1)RW (2)

RW (3)RW (4)

𝛷𝛹

𝛩

(b)

Figure 12: (a) Momentum unloading mode, reaction wheels in pyramid configuration. (b) Closeup.

14 International Journal of Aerospace Engineering

Page 15: Research Article Generic Model of a Satellite Attitude ...

The simulations of the momentum unloading mode con-firmed that the developed control algorithm diminishesangular velocities of reaction wheels in their various configu-rations. The desaturation process has a significant impact onsatellite attitude control and indicates that in the momentumunloading mode, a satellite should not be commanded to per-form tasks which require high accuracy pointing.

5. Conclusions

An attitude control system was developed which fulfils twomain tasks: provides a satellite with sufficient attitude controlcapabilities in the detumbling and normal modes of opera-tion and ensures adequate performance of control actuatorsby the momentum unloading control process. The actuatorsof the control system are magnetorquers and reaction wheels.The software architecture allows scalability of the system tovarious numbers and configurations of actuators as well aschanging PID regulator parameters, adjusting to a particularcase. The model of the attitude control system was integratedwith the satellite simulation software developed at theWarsaw University of Technology. The results of simulationsof the control system operation are satisfying. The modelmay be a useful tool for development of an attitude controlsystem to be operating in real time and embedded into nano-satellite hardware. Also, the simulation results presented inthe paper may be useful for similar system development. Infurther research, the software is intended to be used for othermissions and other control strategies.

Nomenclature

Vectors and matrices: BoldAðtÞ, Bðt, xÞ: Satellite inertia matricesbR: Viscous friction coefficient of the RW

motor shaftBF : Magnetic field vectorBQ: The Earth magnetic field in a

magnetorquer coordinate systemBQ = BxQ ByQ BzQ

� �TCBA: Transformation matrix from

coordinate system A to system BD0: Nominal magnitude of magnetic

dipole generated by onemagnetorquer

DQ: Magnetic dipole generated by themagnetorquer coil, DQ =DQ 0 0

� �TD0: Diagonal matrix composed of mag-

netorquers nominal dipole momentsD0 = diag D01 ⋯ D0NQ

Fð∗Þ: Force generated by (∗)fð∗Þ: Satellite loads fð∗Þ = Fð∗Þ Mð∗Þ

� �TGð∗Þ: Matrix representing actuator

configurationHR: Angular momentum of the reaction

wheel

iR: Armature current of the reactionwheel motor

Ixx, Iyy, Izz : Satellite moments of inertiaIxy, Iyz , Ixz : Satellite products of inertiaIRx : Moment of inertia of the reaction

wheel about its spin axisI: Satellite moments of inertia matrixIRW: Diagonal matrix containing the

moments of inertia of reaction wheelskR: Torque constant of the RW motorKPð∗ÞKDð∗ÞKIð∗Þ: Gain matrices: proportional,

derivative, and integralMð∗Þ: Torquem: Satellite massN : Total number of actuatorsNð∗Þ: Number of devicesqI : Quaternion of a satellite attitude in

the inertial frame qI =q0 q1 q2 q3½ �T

Q: Matrix of angular velocities inattitude equation

rI : Satellite position in the inertialcoordinate system (ECI), rI =xI yI zI½ �T

R: Satellite yaw rateR: Diagonal matrix containing reaction

wheel motor constants, R = diagkR1 ⋯ kRNR

Sx, Sy, Sz : Static mass moments of the satellitet: TimeT : Time constant of the integral

controllerTorb: Satellite orbital perioduð∗Þ: Actuator control signalsvF : Satellite linear velocity in the satellite

body frame, vF = U V W½ �Tx: Satellite state vectorζ: Specified control dumping ratioσ: Attitude error vectorωn: Specified linear control bandwidthωF : Satellite angular velocity in the

satellite body frame, ωF =P Q R½ �T

ωR: Satellite angular velocity in the

actuator coordinate system ωR =ωRx ωRy ωRz

� �TΩR: Spin velocity of a single reaction

wheelO: Satellite linear and angular velocity

matrixΩð∗Þ: Reaction wheel angular velocities.

Indices

(∗): Reflects the device or coordinate systemA: Aerodynamics

15International Journal of Aerospace Engineering

Page 16: Research Article Generic Model of a Satellite Attitude ...

C: Controlcmd: Commandede: Difference between the commanded and current

valuesF: In the satellite body coordinate systemG: GravityM: Magnetic field (acting on the bus)Q: MagnetorquersR: Reaction wheelsSRP: Solar radiation pressureT : Propulsion (not used in this study).

Superscripts

+: Matrix pseudoinverse× : Matrix from vectors of cross-productT : Vector, matrix transposedot: Differentiation in time.

Data Availability

The data described in the article is a result of numerical simu-lation. Although the data itself is not available, a description ofhow it was obtained is provided. The data used to support thefindings of this study are available from the correspondingauthor upon request.

Conflicts of Interest

The authors declare that there is no conflict of interest.

Acknowledgments

This research was supported by the European Commission’sHorizon 2020 Program [grant number 687490].

References

[1] J. Narkiewicz, S. Topczewski, and M. Sochacki, “Nanosatelliteformation flying for maritime applications,” in 7th InternationalScience and Technology Conference on Naval Technologies forDefence and Security NATCON, pp. 345–351, Gdańsk, 2016.

[2] D. T. Gerhardt and S. E. Palo, “Passive magnetic attitude con-trol for CubeSat spacecraft,” in 24th annual AIAA/USU Con-ference on Small Satellites, pp. 1–10, UT, 2010.

[3] J. F. Trégouët, D. Arzelier, D. Peaucelle, C. Pittet, andL. Zaccarian, “Reaction wheels desaturation using magnetor-quers and static input allocation,” IEEE Transactions on Con-trol Systems Technology, vol. 23, no. 2, pp. 525–539, 2015.

[4] E. A. Hogan and H. Schaub, “Three-axis attitude control usingredundant reaction wheels with continuous momentumdumping,” Journal of Guidance, Control, and Dynamics,vol. 38, no. 10, pp. 1865–1871, 2015.

[5] F. Giulietti, A. A. Quarta, and P. Tortora, “Optimal controllaws for momentum-wheel desaturation using magnetor-quers,” Journal of Guidance, Control, and Dynamics, vol. 29,no. 6, pp. 1464–1468, 2006.

[6] X. Chen, W. H. Steyn, S. Hodgart, and Y. Hashida, “Optimalcombined reaction-wheel momentum management forEarth-pointing satellites,” Journal of Guidance, Control, andDynamics, vol. 22, no. 4, pp. 543–550, 1999.

[7] R. Wisniewski, “Linear time-varying approach to satellite atti-tude control using only electromagnetic actuation,” Journal ofGuidance, Control, and Dynamics, vol. 23, no. 4, pp. 640–647,2000.

[8] M. L. Psiaki, “Magnetic torquer attitude control via asymp-totic periodic linear quadratic regulation,” Journal of Guid-ance, Control, and Dynamics, vol. 24, no. 2, pp. 386–394,2001.

[9] H. C. Polat, J. Virgili-llop, and M. Romano, “Survey, statisticalanalysis and classification of launched CubeSat missions withemphasis on the attitude control method,” Journal of SmallSatellites, vol. 5, pp. 513–530, 2016.

[10] Y. K. Chang, B. H. Lee, J. W. Choi, M. Y. Yun, and S. J. Kang,“Rapid initial detumbling strategy for micro/nanosatellitesusing a pitch bias momentum system,” Transactions of theJapan Society for Aeronautical and Space Sciences, vol. 50,no. 167, pp. 63–69, 2007.

[11] W. H. Steyn, Y. Hashida, and V. Lappas, “An attitude controlsystem and commissioning results of the SNAP-1 nanosatel-lite,” in 14th annual AIAA/USU Conference on Small Satellites,pp. 1–8, UT, 2000.

[12] W. Steyn and Y. Hashida, “An attitude control system for alow-cost earth observation satellite with orbit maintenancecapability,” in 13th annual AIAA/USU Conference on SmallSatellites, pp. 1–13, 1999.

[13] Y. Y. Lin and C. T. Wang, “Detumbling of a rigid spacecraft viatorque wheel assisted gyroscopic motion,” Acta Astronautica,vol. 93, pp. 1–12, 2014.

[14] W. Flatley, W. Morgenstern, A. Reth, and F. Bauer, “A B-dotacquisition controller for the RADARSAT spacecraft,” NASA,pp. 79–90, 1996.

[15] G. Avanzini and F. Giulietti, “Magnetic detumbling of a rigidspacecraft,” Journal of Guidance, Control, and Dynamics,vol. 35, no. 4, pp. 1326–1334, 2012.

[16] B. Wie, Space Vehicle Dynamics and Control, Second Edition,American Institute of Aeronautics and Astronautics, Reston,VA, 2008.

[17] Z. Ismail and R. Varatharajoo, “A study of reaction wheel con-figurations for a 3-axis satellite attitude control,” Advances inSpace Research, vol. 45, no. 6, pp. 750–759, 2010.

[18] C.-H. Won, “Comparative study of various control methodsfor attitude control of a LEO satellite,” Aerospace Science andTechnology, vol. 3, no. 5, pp. 323–333, 1999.

[19] J. Li, M. Post, T.Wright, and R. Lee, “Design of attitude controlsystems for CubeSat-class nanosatellite,” Journal of ControlScience and Engineering, vol. 2013, Article ID 657182, 15pages, 2013.

[20] M. Fakhari Mehrjardi, H. Sanusi, M. A. M. Ali, and M. A.Taher, “PD controller for three-axis satellite attitude controlusing discrete Kalman filter,” in 2014 International Conferenceon Computer, Communications, and Control Technology(I4CT), pp. 83–85, Langkawi, Malaysia, 2014.

[21] S. R. Omar, “An inverese dynamics satellite attitude determi-nation and control system with autonomous calibration,” in29th annual AIAA/USU Conference on Small Satellites,pp. 1–12, UT, 2015.

[22] M. Moradi, “Satellite neuro-PD three-axis stabilization basedon three reaction wheels,” Journal of Aerospace Engineering,vol. 27, no. 1, pp. 177–184, 2014.

[23] B. Wie and J. Lu, “Feedback control logic for spacecraft eigen-axis rotations under slew rate and control constraints,” Journal

16 International Journal of Aerospace Engineering

Page 17: Research Article Generic Model of a Satellite Attitude ...

of Guidance, Control, and Dynamics, vol. 18, no. 6, pp. 1372–1379, 1995.

[24] M. J. Sidi, Spacecraft Dynamics and Control, CambridgeUniversity Press, New York, 2014.

[25] P. J. Camillo and F. L. Markley, “Orbit-averaged behavior ofmagnetic control laws for momentum unloading,” Journal ofGuidance, Control, and Dynamics, vol. 3, no. 6, pp. 563–568,1980.

[26] “Sinclair Interplanetary, Nanosatellite reaction wheels (RW-0.03),” January 2018, http://www.sinclairinterplanetary.com/reactionwheels/30mNm-secwheel2016a.pdf?attredirects=0.

[27] “Innovative Solutions In Space, Magnetorquer board(iMTQ),” January 2018, https://www.cubesatshop.com/wp-content/uploads/2016/06/iMTQ-Brochure-v1.pdf.

17International Journal of Aerospace Engineering