Top Banner
Attitude Tracking Control of a Small Satellite in Low Earth Orbit Amit K. Sanyal Zachary Lee-Ho In this paper, we modify and apply a robust almost global attitude tracking control scheme to the model of a small satellite. The control scheme, which has been reported in prior literature, is modified to take into account the actuator constraints and actuator configuration of this satellite, which are based on a small satellite currently being developed at the University of Hawaii. The actuators consist of three magnetic torquers and one small reaction wheel. The mass and inertia properties correspond to the known values for this satellite. The satellite is in circular low earth orbit of altitude 600 km and its dynamics model includes gravity, atmospheric and geomagnetic effects. The control strategy used here achieves almost global asymptotically stable attitude trajectory tracking, which implies that the desired attitude trajectory is tracked from all initial conditions on the state except for those that lie on a zero-volume subset within the state space. The continuous feedback control law is also globally defined. Feedback control gains are continuously varied based on known actuator constraints and tracking errors. The almost global asymptotic tracking property can be shown using a generalized Lyapunov analysis on the nonlinear state space of the attitude dynamics. The control torque obtained from this almost-globally-stabilizing feedback control law is partitioned so that each actuator generates a part of this control torque that is within its saturation limits. The control law for the reaction wheel has a singularity when the reaction wheel axis is perpendicular to the local geomagnetic field. To avoid actuator saturation, the control inputs to the actuators are kept constant whenever any actuator reaches a certain fraction of its saturation value. Numerical simulation results for two de-tumbling maneuvers, one where the control law singularity does not appear and one where it does, confirm that the desired attitude trajectory is tracked almost globally. I. Introduction Small satellites, below 100 kg in mass, are being used in an expanding set of applications because of their shorter development and deployment cycles compared to larger satellites. Errors in orbit insertion, disturbances and unmodeled forces and moments can have a large effect on the attitude dynamics of these satellites due to their small inertia. We consider control of the attitude dynamics of a small satellite in low earth orbit under the effects of gravity, atmospheric drag and geomagnetic moments. The hardware configuration and constraints of the attitude control actuators are considered in designing the control law. The control task is for the satellite to track desired attitude maneuvers that have been uplinked from the ground or maneuvers that have to be carried out to recover from faults. Attitude control of a rigid body is a benchmark problem in nonlinear control with applications to aerial, ground and underwater vehicles, besides spacecraft. This problem has been studied under various assump- tions and scenarios; as a small sample of the literature shows. 1–8 However, much of the prior research on attitude control and stabilization of a rigid body has been carried out using minimal three-coordinate or quaternion representations of the attitude. Any minimal representation necessitates a local analysis since such representations have kinematic singularities and cannot represent attitude globally. The quaternion representation is ambiguous, since for every possible attitude there are two sets of unit quaternion repre- sentations. Since unit quaternions distinguish between principal angle rotations of 0 and 2π, this leads to continuous controllers that rotate the spacecraft needlessly from principal angle 2π to 0, resulting in lack of stability of the desired equilibrium or trajectory. This was termed the unwinding phenomenon by Bhat and Assistant Professor, Mechanical Engineering, University of Hawaii at Manoa, Honolulu, HI 96822, [email protected] MS student, Mechanical Engineering, University of Hawaii at Manoa, Honolulu, HI 96822, [email protected] 1 of 14 American Institute of Aeronautics and Astronautics AIAA Guidance, Navigation, and Control Conference 10 - 13 August 2009, Chicago, Illinois AIAA 2009-5902 Copyright © 2009 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved.
14

Attitude Tracking Control of a Small Satellite in Low ...mae2.nmsu.edu/~asanyal/Sanyal_res/AIAA-2009-5902-938.pdf · Attitude Tracking Control of a Small Satellite in Low ... The

Apr 01, 2019

Download

Documents

hacong
Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Attitude Tracking Control of a Small Satellite in Low ...mae2.nmsu.edu/~asanyal/Sanyal_res/AIAA-2009-5902-938.pdf · Attitude Tracking Control of a Small Satellite in Low ... The

Attitude Tracking Control of a Small Satellite in Low

Earth Orbit

Amit K. Sanyal∗

Zachary Lee-Ho†

In this paper, we modify and apply a robust almost global attitude tracking control

scheme to the model of a small satellite. The control scheme, which has been reported

in prior literature, is modified to take into account the actuator constraints and actuator

configuration of this satellite, which are based on a small satellite currently being developed

at the University of Hawaii. The actuators consist of three magnetic torquers and one small

reaction wheel. The mass and inertia properties correspond to the known values for this

satellite. The satellite is in circular low earth orbit of altitude 600 km and its dynamics

model includes gravity, atmospheric and geomagnetic effects. The control strategy used

here achieves almost global asymptotically stable attitude trajectory tracking, which implies

that the desired attitude trajectory is tracked from all initial conditions on the state except

for those that lie on a zero-volume subset within the state space. The continuous feedback

control law is also globally defined. Feedback control gains are continuously varied based

on known actuator constraints and tracking errors. The almost global asymptotic tracking

property can be shown using a generalized Lyapunov analysis on the nonlinear state space

of the attitude dynamics. The control torque obtained from this almost-globally-stabilizing

feedback control law is partitioned so that each actuator generates a part of this control

torque that is within its saturation limits. The control law for the reaction wheel has a

singularity when the reaction wheel axis is perpendicular to the local geomagnetic field. To

avoid actuator saturation, the control inputs to the actuators are kept constant whenever

any actuator reaches a certain fraction of its saturation value. Numerical simulation results

for two de-tumbling maneuvers, one where the control law singularity does not appear and

one where it does, confirm that the desired attitude trajectory is tracked almost globally.

I. Introduction

Small satellites, below 100 kg in mass, are being used in an expanding set of applications because oftheir shorter development and deployment cycles compared to larger satellites. Errors in orbit insertion,disturbances and unmodeled forces and moments can have a large effect on the attitude dynamics of thesesatellites due to their small inertia. We consider control of the attitude dynamics of a small satellite inlow earth orbit under the effects of gravity, atmospheric drag and geomagnetic moments. The hardwareconfiguration and constraints of the attitude control actuators are considered in designing the control law.The control task is for the satellite to track desired attitude maneuvers that have been uplinked from theground or maneuvers that have to be carried out to recover from faults.

Attitude control of a rigid body is a benchmark problem in nonlinear control with applications to aerial,ground and underwater vehicles, besides spacecraft. This problem has been studied under various assump-tions and scenarios; as a small sample of the literature shows.1–8 However, much of the prior research onattitude control and stabilization of a rigid body has been carried out using minimal three-coordinate orquaternion representations of the attitude. Any minimal representation necessitates a local analysis sincesuch representations have kinematic singularities and cannot represent attitude globally. The quaternionrepresentation is ambiguous, since for every possible attitude there are two sets of unit quaternion repre-sentations. Since unit quaternions distinguish between principal angle rotations of 0 and 2π, this leads tocontinuous controllers that rotate the spacecraft needlessly from principal angle 2π to 0, resulting in lack ofstability of the desired equilibrium or trajectory. This was termed the unwinding phenomenon by Bhat and

∗Assistant Professor, Mechanical Engineering, University of Hawaii at Manoa, Honolulu, HI 96822, [email protected]†MS student, Mechanical Engineering, University of Hawaii at Manoa, Honolulu, HI 96822, [email protected]

1 of 14

American Institute of Aeronautics and Astronautics

AIAA Guidance, Navigation, and Control Conference10 - 13 August 2009, Chicago, Illinois

AIAA 2009-5902

Copyright © 2009 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved.

Page 2: Attitude Tracking Control of a Small Satellite in Low ...mae2.nmsu.edu/~asanyal/Sanyal_res/AIAA-2009-5902-938.pdf · Attitude Tracking Control of a Small Satellite in Low ... The

Bernstein.2 To avoid this phenomenon, a discontinuous quaternion-based controller that overcomes unwind-ing is used.9 Nevertheless, discontinuous dynamics entail special difficulties,10 and may lead to chattering inthe vicinity of a discontinuity, especially in the presence of sensor noise or disturbances. It is thus of interestto determine which closed-loop properties can be achieved under continuous control.

Spacecraft attitude control is an inherently strongly nonlinear (in particular, quadratic) problem whoseconfiguration space is nonlinear: the special orthogonal group SO(3) of 3 × 3 rotation matrices. Thereis very little natural damping on spacecraft attitude motion. Although linear controllers can be used formaneuvers over small angles, the desire for robust, minimum-fuel or minimum-time operation suggests thatcontrol systems that are tuned for operation on SO(3) are advantageous for large-angle maneuvers.11,12However, the compactness of SO(3) presents difficulties with regard to global asymptotic stabilization, thatis, Lyapunov stability of a desired equilibrium along with global convergence. As shown by Koditschek,13 theminimum number of equilibria for a Lyapunov-like function on SO(3) is 4. The mere existence of multipleequilibria precludes global asymptotic stability of the desired equilibrium in the physical configuration space.

The largest domain of stability that can be obtained with continuous feedback control laws for attitudestabilization and tracking is almost global.14–16 By almost global stabilization, we mean that all trajectoriesexcept those beginning in a set of zero volume measure in the state space, converge to the desired equilibriumstate (for stabilization) or the desired state trajectory (for tracking). In practice, this property is effectivelyequivalent to global stabilization since no real trajectory can remain on a set of measure zero. The first paperconstructed a Lyapunov-like function obtained from energy-shaping ideas13 on the nonlinear state space ofattitude motion, to obtain almost global stabilization of a desired attitude equilibrium for a rigid satellite inorbit. It also included the effects of actuator saturation in the analysis to obtain a continuous almost globallystabilizing control law that satisfies saturation constraints. The second paper, which appears in last year’sGuidance, Navigation and Control conference, extends this stabilization scheme to track a desired attitudemaneuver in the presence of a disturbance moment generated by atmospheric drag. The resulting controllaw has a discontinuity when the angular velocity tracking error vanished. The third paper generalizes theseresults to an adpative control scheme that is robust to imperfect knowlegde of inertia and disturbancescreated by a linear internal model. These disturbances have known frequency content but unknown boundedamplitudes. The third paper makes this tracking control scheme adaptive for application to spacecrafttracking control with unknown inertia properties and a disturbance torque with known frequency contentbut unknown amplitude. However, these tracking control schemes15,16 do not take into account actuatorconstraints for any particular actuator configuration used.

The motivation for this paper is to apply these recent results14–16 to a model of a small spacecraft withknown actuator saturation limits. This is a small satellite in a low-Earth orbit satellite that has one reactionwheel and three magnetic torquers as actuators. A brief outline of this paper is given here. In Section II,the model of the small satellite being considered is presented, along with its actuator configuration. Anoverall tracking control law which asymptotically tracks the desired trajectory almost globally is presentedin Section III. This section also gives a scheme for feedback gain sizing of the attitude and angular velocityfeedback gains based on attitude and angular velocity tracking errors, so that the required control torqueis well within the saturation constraints of the actuators. The almost global asymptotic properties of thiscontrol law have been shown analytically and numerically in related earlier work.15,16 In Section IV, wepresent the generation and partitioning of the required control torque by the actuators. The control law forthe reaction wheel is obtained first, and this control law has a singularity that depends on the alignmentof the local geomagnetic field in orbit in the satellite body frame. The difference between the total torquerequired for the maneuver and the reaction wheel torque is used to obtain the control law for the magnetictorquers. To avoid saturation of the reaction wheel when the satellite is close to the singularity configurationfor the reaction wheel control law, the control input to all actuators is kept constant when any actuatorreaches a certain fraction of its saturation level. The reaction wheel control law demands substantially lesstorque when the satellite moves out of this singular configuration; in the satellite’s natural orbital motion,this singularity occurs only instantaneously. Simulation results for two maneuvers that de-tumble the smallsatellite and align its attitude to the LVLH frame in low earth circular orbit are presented in Section V. Oneof these maneuvers does not reach a singularity configuration, while the other maneuver passes through asingularity. Results for both these maneuvers demonstrate the practically globally asymptotic and robustproperties of this attitude and angular velocity tracking control algorithm while maintaining actuator torqueand momentum constraints. The concluding Section VI has a summary of results presented, and presentspossible future developments and applications.

2 of 14

American Institute of Aeronautics and Astronautics

Page 3: Attitude Tracking Control of a Small Satellite in Low ...mae2.nmsu.edu/~asanyal/Sanyal_res/AIAA-2009-5902-938.pdf · Attitude Tracking Control of a Small Satellite in Low ... The

II. Model for Spacecraft Attitude Dynamics and Control

We define the attitude of the spacecraft as the relative orientation of a body-fixed coordinate frame toan inertial frame, which in the case of the satellite we consider, is geocentric. The coordinate axes of theseframes are related by a proper orthogonal matrix, called the rotation matrix, which gives a global and uniquerepresentation of the attitude. The attitude can also be represented by coordinate sets, like the Euler angles,quaternions, or Rodrigues parameters.17,18 We represent the attitude globally by using rotation matrices.These matrices form a group under matrix multiplication, denoted by SO(3), where

SO(3) = R ∈ R3×3 : RTR = I = RRT, det(R) = 1,

where I denotes the 3× 3 identity matrix. We denote the attitude of the spacecraft modeled as a rigid bodyby R ∈ SO(3). Note that we describe rotation matrices as members of the 3-dimensional quadratic Lie groupSO(3) and not the 9-dimensional linear matrix space R3×3. This point is reiterated in Section V, where wepresent numerical simulation results obtained using a Lie Group Variational Integrator, which is a numericalalgorithm obtained from discrete variational mechanics on the nonlinear state space TSO(3).19

II.A. Satellite Model for Attitude Control

The model of the small satellite we use here is based on the satellite being developed for the first plannedmission of the Hawaii Space Flight Laboratory (HSFL) at the University of Hawaii. This satellite is about50 kg in mass. The principal moment of inertia matrix for this satellite is

J = diag(2.13, 3.04, 3.09) kg·m2.

The Attitude Determination and Control Subsystem (ADCS) of the satellite consists of actuators, sensors,data acquisition boards and an onboard processor that runs the ADCS control and estimation codes. Theattitude actuators consist of three magnetic torquers and one small reaction wheel, as depicted in Figure 1of the satellite with side panels removed. The satellite is shaped like an octagonal cylinder, with the sidepanels to hold solar cell arrays.

Figure 1. View of the HSFL small satellite with attitude actuators

The magnetic torquers have constraints on their inputs, which are the magnetic moments of the torquercoils. The reaction wheel has constraints on its stored angular momentum and its motor torque. Forthe HSFL small satellite, the magnetic moments produced by the torquers have a maximum of 25 A·m2.

3 of 14

American Institute of Aeronautics and Astronautics

Page 4: Attitude Tracking Control of a Small Satellite in Low ...mae2.nmsu.edu/~asanyal/Sanyal_res/AIAA-2009-5902-938.pdf · Attitude Tracking Control of a Small Satellite in Low ... The

The reaction wheel on this satellite is designed to produce a nominal torque of around 2 milli-N·m with amaximum torque of 3 milli-N·m and maximum angular momentum of 30 milli-N·m·s. These constraints areused in designing the control gains for our attitude feedback control scheme when carrying out de-tumblingand slew maneuvers. We choose the satellite body frame such that the body X-axis lines up with the centerof one of the panels, and the Z-axis points to the “bottom” surface which has antennas and an imager; thissurface is supposed to point towards the Earth in orbit. The center shelf in the structure has most of thebattery power sources and avionics mounted on it, including an inertial measurement unit with rate gyrosand a magnetometer. Not shown in Figure 1 are two sun sensors that will be mounted on two of the sidepanels of the structure.

The axis of the reaction wheel is aligned along the body X-axis, and its moment of inertia in the bodyframe is given by

Iw = diag(1.47, 1.26, 1.26)× 10−5 kg·m2.

The three magnetic torquers are fixed to the structure along directions given by the unit vectors

u1 =1√3

11−1

, u2 =1√3

1−11

, u3 =

001

,

in the body frame. The satellite attitude motion is fully actuated with the control scheme outlined here.

II.B. Equations of Motion for Attitude Dynamics

We consider the satellite’s attitude dynamics in the presence of a gravity-gradient potential that is determinedby the attitude, an atmospheric drag moment about its center of mass, and control moments due to magnetictorquers and a reaction wheel. Let Ω ∈ R3 be the angular velocity of the satellite measured in the bodyframe. The attitude kinematics is given by Poisson’s equation

R = RΩ×, (1)

where (·)× : R3 → so(3) is the skew-symmetric mapping given by

Ω× =

Ω1

Ω2

Ω3

×

=

0 −Ω3 Ω2

Ω3 0 −Ω1

−Ω2 Ω1 0

.

We denote the space of 3× 3 real skew-symmetric matrices by so(3); this is the Lie algebra of the Lie groupSO(3). Note that (·)× is also the cross-product in R3: x×y = x× y.

Let U : SO(3) → R be a gravity potential dependent on the attitude and let J denote the satellite’smoment of inertia matrix in the body frame. The attitude dynamics of the satellite is given by the followingequations of motion

JΩ + Ω× JΩ = Mg(R) + Mad(R,Ω) + τ, (2)

where τ is the control input (torque) vector, Mad(R,Ω) is the sum of known nonconservative moments actingon the system like atmospheric drag, and Mg(R) is the moment due to the gravity potential U(R):

M×g (R) =

∂U

∂R

T

R−RT ∂U

∂R, R =

r1

r2

r3

∈ SO(3), or

Mg(R) = r1 × vr1 + r2 × vr2 + r3 × vr3 ,∂U

∂R=

vr1

vr2

vr3

.

The partial derivative ∂U∂R ∈ R3×3 is defined such that

∂U∂R

ij= ∂U

∂Rij. The state space for the spacecraft

attitude dynamics is thus the tangent bundle TSO(3). A similar dynamics model, without control torques,has been used to study the attitude estimation problem.19

4 of 14

American Institute of Aeronautics and Astronautics

Page 5: Attitude Tracking Control of a Small Satellite in Low ...mae2.nmsu.edu/~asanyal/Sanyal_res/AIAA-2009-5902-938.pdf · Attitude Tracking Control of a Small Satellite in Low ... The

II.C. Trajectory Tracking Problem Formulation

In this part of this section, we introduce the attitude and angular velocity trajectory tracking problem, andformulate it in terms of tracking errors in attitude and angular velocity. We specify the desired trajectoryby the initial attitude Rd(0) and the angular velocity as a function of time Ωd(t), for time t ≥ 0. Further,Ωd(t) and Ωd(t) are bounded, so that the attitude rate of change is given by

Rd(t) = Rd(t)Ωd(t)×, given Rd(0),Ωd(t). (3)

This desired trajectory may be obtained from an open-loop motion planning or trajectory generation scheme,like an open-loop optimal or sub-optimal control scheme.

We first define the attitude and angular velocity tracking errors as follows:

Q(t) RTd (t)R(t) = exp

ζ(t)×

, ω(t) Ω(t)−QT(t)Ωd(t), (4)

where ζ is the product of the principal angle with the unit vector along the principal axis representing therotation Q. These definitions and equation (3) lead to the attitude error kinematics equation:

Q = QΩ−QTΩd

× = Qω×. (5)

Note that the attitude error kinematics is also “left-invariant” like the original attitude kinematics (1), i.e.,if Ql = CQ where C ∈ SO(3) is constant, then Ql = Qlω×.

Now we re-write the attitude dynamics equation (2) by susbstituting for Ω(t) = ω(t) + QT(t)Ωd(t) from(4). We also evaluate the gravity potential U(R) as a function of the attitude error Q(t) after substitutingR(t) = Rd(t)Q(t) from (4). Therefore, we obtain:

J ω = Jω×QTΩd −QTΩd

ω + QTΩd

×Jω + QTΩd

+ Mg(RdQ) + Mad(R,Ω) + τ. (6)

The trajectory tracking error kinematics and dynamics (5) and (6) depend on Q, ω, Ωd, Ωd, the momentdue to the potential Mg = Mg(RdQ), the atmospheric drag moment Mad(R,Ω) and control input τ .

III. Attitude and Angular Velocity Feedback Tracking Control

In this section, we give a control law that achieves the control task of asymptotically tracking thedesired attitude and angular velocity as specified by (3), following the kinematics and dynamics (5)-(6).The controller obtained achieves almost global asymptotic tracking, i.e., the attitude and angular velocityconverge to the desired trajectory from all initial states excluding a set that is open and dense in the tangentbundle TSO(3) with zero volume measure.

III.A. Feedback Control Law

Let Φ : R+ → R+ be a C2 function that satisfies Φ(0) = 0 and Φ(x) > 0 for all x ∈ R+. Furthermore, letΦ(·) ≤ α(·) where α(·) is a Class-K function.20 Let K > 0 be the attitude feedback control gain matrix,with K = diag(k1, k2, k3) such that 0 < k1 < k2 < k3. Therefore Φ(trace(K − KQ)) is a Morse functionon SO(3) for Q ∈ SO(3), and its critical points are non-degenerate and hence isolated, according to theMorse lemma.21 On a nonlinear configuration space like SO(3), Morse functions are suitable candidates forLyapunov-like functions. As shown in prior literature,13–16 such a function has 4 critical points on SO(3),which happens to be the minimum number of critical points that a Lyapunov-like function on SO(3) canhave. Along the kinematics (5), the time derivative of this function is:14

ddt

Φ(trace(K −KQ))

= −Φ(trace(K −KQ))trace(KQω×)

= −Φ(trace(K −KQ))ωTk1e

×1 QTe1 + k2e

×2 QTe2 + k3e

×3 QTe3

. (7)

The global minimum for this function on SO(3) is given by Q = I, the identity matrix.

5 of 14

American Institute of Aeronautics and Astronautics

Page 6: Attitude Tracking Control of a Small Satellite in Low ...mae2.nmsu.edu/~asanyal/Sanyal_res/AIAA-2009-5902-938.pdf · Attitude Tracking Control of a Small Satellite in Low ... The

We propose the following control law to asymptotically track the desired attitude and angular velocity:

τ = −Lω + JQTΩd +QTΩd

×JQTΩd + Φ(trace(K −KQ))

k1e

×1 QTe1 + k2e

×2 QTe2

+ k3e×3 QTe3

−Mg(RdQ)−Mad(RdQ,Ω).

(8)

In this control law, the matrix L is a positive definite angular velocity feedback gain matrix. Note thatthis control law, and hence the trajectories of the closed-loop system, are continuous with respect to thetrajectory tracking error variables Q and ω. This eliminates chattering, which may occur when trackingcontrol laws that are discontinuous across a sliding surface are used. Previous results15 have shown that(Q,ω) = (I, 0) is a stable equilibrium of the closed loop system consisting of (5)-(6) and (8). Further, it hasbeen shown15,16 this equilibrium is almost globally asymptotically stable. By “almost globally” we meanthat the domain of attraction is the whole state space (which is the tangent bundle TSO(3)), except for asubset of zero volume measure. In practice, this is as good as global asymptotic trajectory tracking, sinceno real trajectory can remain in a zero volume subset of the state space for indefinite time.

III.B. Gain Sizing to meet Actuator Constraints

The control law (8) leads to almost global asymptotic tracking of an attitude and angular velocity trajectoryfor any positive diagonal matrix K with distinct entries and any positive definite matrix L. However, if theattitude actuators have known constraints and if the desired trajectory does not demand control torques thatthe actuators cannot deliver, then these gains could be continuously adjusted so that the actuator constraintsare satisfied. Here we modify this basic scheme by modifying the control gains K and L so that actuatorsaturation bounds are satisifed for slewing and de-tumbling maneuvers. We consider slewing maneuvers asthose in which the satellite has to maintain an inertial attitude with zero angular velocity. For a satellite inorbit, we consider a de-tumbling maneuver as one in which the final attitude is constant with respect to thelocal orbital (LVLH) frame. The gains are continuously varied based on state feedback so that the knownactuator bounds for the satellite model in Section II are satisfied.

The gain matrices are modified by multiplying them with positive scalar feedback gain factors as follows:

K = k(Q)K0, L = l(ω)L0, (9)

where K0 is a positive diagonal matrix with distinct entries and L0 is a positive definite matrix. Althoughdifferent schemes may be used to vary these gain factors, we choose to vary the scalar gain factors k(Q)and l(ω) according to the initial tracking errors. If Q = exp(ζ×), then ζ denotes the principal angle valuecorresponding to Q. The gain factors are varied according to

k(Q(t)) =κ0(a1 + ζ(t))

a2 + ζ(0) ,

l(ω(t)) =λ0(c1 + ω(t))

c2 + ω(0) ,

(10)

where κ0, λ0, a1, a2, c1 and c2 are constant positive scalars. This choice of feedback gains is made keepingin mind the Lyapunov analysis given in Sanyal and Chaturvedi15 where the Lyapunov function

V (Q, ω) =12ωTJω + Φ

trace(K −KQ)

. (11)

was used to show the almost global tracking performance of the control law (8). Since the value of thisfunction decreases for the feedback system, one can adjust the values of κ0, λ0, a1, a2, c1 and c2 such thatthe maximum control torque using (8) is obtained initially (at time t = 0). We then need to ensure that thiscontrol torque is well within the actuator constraints.

IV. Generation of Control Torque by Actuators

The control law presented in the last section provides almost global asymptotic tracking of the desiredattitude and angular velocity trajectory, assuming that actuator saturation bounds are not exceeded by thecontrol torque generated according to (8). In this section, we describe how the required control torque givenby this control scheme is partitioned among the actuators used in this satellite.

6 of 14

American Institute of Aeronautics and Astronautics

Page 7: Attitude Tracking Control of a Small Satellite in Low ...mae2.nmsu.edu/~asanyal/Sanyal_res/AIAA-2009-5902-938.pdf · Attitude Tracking Control of a Small Satellite in Low ... The

IV.A. Torque Generation by Reaction Wheel

A reaction wheel or momentum wheel is an internal (shape) actuator that contributes to the total spacecraftangular momentum, i.e., the angular momentum of the satellite plus reaction wheel. If the reaction wheel spinaxis is aligned with the satellite body X-axis, then the column vector representation of the total spacecraftangular momentum is as follows:

Π = JΩ + lwe1, lw = Iwωw, (12)where ωw is the (scalar) angular speed of the wheel about its axis, and where

e1 =

100

, e2 =

010

, and e3 =

001

are the standard basis column vectors corresponding to the body X-, Y- and Z-axes.The total control torque applied to the satellite can be separated into a part generated by the reaction

wheel τRW and a part generated by the magnetic torquers τMT :

τ = τRW + τMT .

We can thus re-write Euler’s equation (2) for this satellite as follows:

Π + Ω×Π = Mg(R) + Mad(R,Ω) + τMT . (13)

The left-hand side of equation (13) can be expanded on substitution of (12) as follows:

JΩ + lwe1 + Ω× (JΩ + lwe1)

=JΩ + Ω× JΩ + (lwe1 + lwΩ× e1).

Substituting the above expression back onto the left-hand side of equation (13) and comparing with equation(2), we obtain the following expression for the torque generated by the reaction wheel:

τRW = −(lwe1 + lwΩ× e1) =

−lw00

0

Ω3lw−Ω2lw

=

−lw−Ω3lwΩ2lw

. (14)

IV.B. Torque Generation by Magnetic Torquers

The magnetic torquers generate torque by interaction with the local geomagnetic field in Earth orbit. LetuBg denote the unit vector along the local geomagnetic field and Bg denote the scalar strength of this field(with SI unit Tesla). This direction vector may be obtained from a magnetometer reading for the satellitein orbit. Let u represent the unit vector along the magnetic torquer axis and µ denote the scalar magneticmoment (SI unit of A·m2) produced by the torquer. The torque generated by this magnetic torquer is thengiven by the vector

τ0MT = (µBg)u×uBg (15)

in the body frame. Since we have three magnetic torquers, we denote their magnetic moments by µ1, µ2

and µ3. Therefore the total torque generated by the three magnetic torquers as configured in the satellitedescribed in Section II is given by:

τMT = Bg

µ1u

×1 uBg + µ2u

×2 uBg + µ3u

×3 uBg

. (16)

We defineMB [u×1 uBg u×2 uBg u×3 uBg ], (17)

so that the total torque generated by the magnetic torquers can be expressed as:

τMT = BgMBm, where m

µ1

µ2

µ3

. (18)

Clearly, the magnetic torquers cannot generate a torque vector along the direction of the local geomagneticfield and a magnetic torquer cannot generate any torque at all if its axis is aligned with the local geomagneticfield. Therefore full attitude actuation cannot be achieved by magnetic torquers alone.

7 of 14

American Institute of Aeronautics and Astronautics

Page 8: Attitude Tracking Control of a Small Satellite in Low ...mae2.nmsu.edu/~asanyal/Sanyal_res/AIAA-2009-5902-938.pdf · Attitude Tracking Control of a Small Satellite in Low ... The

IV.C. Partitioning of Control Torque among Actuators

Since the magnetic torquers cannot generate any torque component along the geomagnetic field direction,the reaction wheel is used to generate this torque component. The component of the control torque obtainedfrom the control law (8) along the geomagnetic field, is equated with the component of the reaction wheeltorque (14) along the geomagnetic field. Therefore,

τTRW uBg = −lwuBg1 − Ω3lwuBg2 + Ω2lwuBg3 = τTuBg, (19)

where uBg = [uBg1 uBg2 uBg3]T in components. Equation (19) can be used to obtain the rate of the wheelmomentum (wheel motor torque input) at any instant of time, as follows:

lw =−1

uBg1

τTuBg + Ω3lwuBg2 − Ω2lwuBg3

, (20)

with required control torque τ , the local geomagnetic field direction uBg , the wheel momentum lw, and thespacecraft angular velocity Ω known. A singularity arises in calculating lw from equation (20) when thelocal geomagnetic field is perpendicular to the reaction wheel axis (which is the spacecraft’s first body axis),in which case uBg1 = 0. If the nominal desired attitude is to have the satellite body axes aligned with thelocal vertical local horizontal (LVLH) frame axes, this control singularity is likely to occur near the Earth’smagnetic north and south poles. However, this singularity is expected to be instantaneous or to last for avery short duration. Therefore, all actuator inputs will be kept constant whenever any actuator reaches apreset fraction of its saturation level, till the actuator input demanded is below this preset level.

For a satellite with no bias momentum, the angular momentum of the reaction wheel is initially zero, i.e.,lw(0) = 0, and equation (20) could be used to obtain lw(t) for t > 0. The torque generated by the reactionwheel is then obtained from equation (14). Therefore, the total torque generated by the magnetic torquersis given by

τMT = τ − τRW .

Now we are ready to calculate the magnetic torquer inputs (their magnetic moments), which are given bythe vector m defined in equation (18). As a first step, we define two orthonormal vectors that span theplane that is orthogonal to the local geomagnetic field direction, uBg ; this is the plane on which the torquesgenerated by the magnetic torquers lie. These vectors are defined as follows:

nBg τ − τTuBg

τ − τTuBg, bBg u×Bg

nBg . (21)

Nest we define the 3× 2 matrices

Np [nBg bBg ] and A = MTBNp, (22)

where MB is defined by (17). The three rows of A give the components of the direction vectors of the torquegenerated by the three magnetic torquers along the orthonormal vectors nBg and bBg . To obtain the vectorm, we modify equation (18) as follows:

BgNTp MBm = NT

p τMT

or BgATm = NTp τMT . (23)

Note that at any given instant, the matrix A is guaranteed to be full column rank since at least two ofthe magnetic torquers are not aligned with the local geomagnetic field. Therefore, the vector of magneticmoments to be input to the torquers is given by:

m =1

BgA(ATA)−1NT

p τMT . (24)

The calculated magnetic moments can be made to satisfy the torquer input constraints by using the controlgain sizing scheme given in Section III.

8 of 14

American Institute of Aeronautics and Astronautics

Page 9: Attitude Tracking Control of a Small Satellite in Low ...mae2.nmsu.edu/~asanyal/Sanyal_res/AIAA-2009-5902-938.pdf · Attitude Tracking Control of a Small Satellite in Low ... The

V. Simulation Results

The numerical simulation results shown here are carried out using a Lie group variational integrator,which is modified from that presented in some of our earlier results,19,22,23 and includes the effect of non-conservative control and other external torques. Besides being accurate in propagating the dynamics, thisstructured Lie group variational integrator also maintains the orthogonality constraint for the rigid rotationmatrix without need for reprojection from R3×3 to SO(3). The two sets of simulation results follow.

V.A. Satellite Orbit and Environment

In this section, we obtain numerical simulation results for the Hawaii Space Flight laboratory 50 kg smallsatellite in a circular Keplerian Low Earth orbit (LEO) at an altitude of 600 km (orbital rate of 0.001085rad/sec). The satellite orbit has an inclination angle of about 98 in the geocentric inertial frame; thiscorresponds to a sun-synchronous orbit. The actuator configuration for this satellite was described in SectionII. The control scheme has been outlined in Section III. For ease in numerical simulation, we assume thatthe mass distribution in the satellite is such that the principal body frame coincides with the body framedescribed in Section II. This is a good assumption in practice, since the actual principal axes differ fromthe chosen body axes by less than 7 for this satellite. The desired attitude trajectory is the “LVLH hold”trajectory, which aligns the satellite body axes along the LVLH axes, with the body X-axis pointing in theorbit tangential (velocity) direction. In this mode, the satellite’s Z-axis (nadir face) points towards Earth,i.e., towards the center of the orbit. This is the nominal attitude for this satellite, which has an imager andantennas on the nadir-pointing face.

Numerical implementation of the control scheme given in this paper on the spacecraft model is used toverify that the actuator configuration used in the Attitude Determination and Control Subsystem (ADCS) ofthe satellite ensures that the desired tracking performance is achieved. To test the robustness of this controlscheme, we do not use any model of the atmospheric drag moment in the control law, i.e., Mad(R,Ω) = 0in (8). However, we include an atmospheric drag moment inside the dynamics equation (2); this dragmoment is calculated using the NRL-MSISE-00 atmospheric model.24 For obtaining the geomagnetic fieldat a location in orbit, we use the World Magnetic Model.25 Both these models are implemented in theMATLAB Aerospace Control Toolbox. The atmospheric drag moments are calculated to be of the order of10−5 N·m during an orbit. For the simulation results shown here, we assume that the satellite is initiallycrossing the equator northwards.

V.B. Attitude Maneuver without Control Singularity

For the first attitude and angular velocity tracking maneuver considered here, we assume the the initialattitude and initial angular velocity for the satellite are:

R(0) =

0.3333 0.2440 0.9107−0.1149 −0.9482 0.29610.9358 −0.2033 −0.2880

, Ω(0) =

0.0400−0.07890.0400

rad/sec.

These correspond to an initial attitude tracking error corresponding to a principal angle rotation of 90about the principal axis 1√

3[−1 1 − 1]T, and an initial angular velocity tracking error of 0.098√

(6)[1 − 2 1]T

rad/s (about 6/sec) respectively, based on the desired “LVLH hold” attitude trajectory. Using the controlscheme given, we adjust the feedback angular velocity and attitude tracking gains according to equations(9)-(10) such that the desired actuator constraints are met.

The initial tracking errors, obtained from the initial conditions, are used in the control scheme to setthe attitude and angular velocity feedback tracking gains according to equations (8)-(10) with κ0 = 0.098,λ0 = 1.47, a1 = 3.5, a2 = 35, c1 = 2.1 and c2 = 21. We simulate the satellite’s attitude dynamics in feedbackloop for one-fifteenth of an orbit. The results of the simulations can be seen in the plots to follow. The firstset of plots (Figure 2) show that the attitude tracking error (expressed as a principal angle) and the angularvelocity tracking error go to zero when the control scheme is applied to the system. The plots validate theeffectiveness of the tracking control scheme which commands the satellite to orient itself and maintain thedesired “LVLH hold” attitude and angular velocity. The moments generated from each actuator, magnetictorquer rods and reaction wheel respectively can be seen in the second set of plots (Figure 3). These

9 of 14

American Institute of Aeronautics and Astronautics

Page 10: Attitude Tracking Control of a Small Satellite in Low ...mae2.nmsu.edu/~asanyal/Sanyal_res/AIAA-2009-5902-938.pdf · Attitude Tracking Control of a Small Satellite in Low ... The

Figure 2. Attitude tracking error in principal angle (left) and norm of angular velocity tracking error (right) for the

de-tumbling and “LVLH hold” maneuver without any singularities

two plots are significant in determining if the actuators can generate the needed torque within the givenhardware constraints. Through our numerical simulations it can be seen that the desired orientation tothe LVLH frame can be achieved with a reaction wheel momentum less than 3.5 milli-N·m·s in magnitudeand less than 15 A·m2 of magnetic moment from each of the magnetic torquer rods. These values are wellwithin the capabilities of the attitude actuators of the HSFL small satellite, which has a reaction wheelmomentum saturation level of 30 milli-N·m·s and magnetic torquer saturation limit of 25 A·m2. The totaltorque produced is also well within the capabilities of these actuators. The last plot (Figure 4) displays

Figure 3. Magnetic Torquer moments (left) and reaction wheel momentum (right) required for the satellite de-tumbling

and “LVLH hold” maneuver without any singularities

the total required torque magnitude from all actuators for this maneuver, as well as the orientation of thegeomagnetic field relative to the satellite during this maneuver. Identification of magnetic field vectors inthe satellite body frame assist us in the placement of the reaction wheel and knowing when the reactionwheel control scheme given by equation (20) is going to be singular. In this maneuver, the satellite attitudenever reaches close to a singularity configuration for the reaction wheel control law (20), and therefore theactuators never approach close to their saturation limits. These results show that satisfactory de-tumblingand tracking of the LVLH hold attitude is achieved in less than one-thirtieth of an orbit.

10 of 14

American Institute of Aeronautics and Astronautics

Page 11: Attitude Tracking Control of a Small Satellite in Low ...mae2.nmsu.edu/~asanyal/Sanyal_res/AIAA-2009-5902-938.pdf · Attitude Tracking Control of a Small Satellite in Low ... The

Figure 4. Total control torque magnitude generated by the reaction wheel and magnetic torquer rods (left) and unit

vector components giving the direction of the local geomagnetic field in body frame (right), during the de-tumbling

and “LVLH hold” maneuver without any singularities

V.C. Attitude Maneuver with Control Singularity

For the second attitude and angular velocity tracking maneuver considered here, we assume the the initialattitude and initial angular velocity for the satellite are:

R(0) =

0.0000 0.7071 0.7071−0.6018 −0.5647 0.56470.7986 −0.4255 0.4255

, Ω(0) =

0.0314−0.06290.0304

rad/sec.

These correspond to an initial attitude tracking error corresponding to a principal angle rotation of 90about the principal axis 1√

2[0 1 − 1]T, and an initial angular velocity tracking error of 0.077√

(6)[1 − 2 1]T rad/s

(about 4.4/sec) respectively, based on the desired “LVLH hold” attitude trajectory. These initial conditionsare such that the satellite attitude goes through an instantaneous singularity configuration for the reactionwheel control law (20). The feedback angular velocity and attitude tracking gains as given by equations(9)-(10) are set identical to those used in the earlier singularity-free maneuver.

Figure 5. Attitude tracking error in principal angle (left) and norm of angular velocity tracking error (right) for the

de-tumbling and “LVLH hold” maneuver with a control law singularity

We simulate the satellite’s attitude dynamics in feedback loop for one-fifteenth of an orbit beginning witha northwards Equator crossing. The results of the simulations can be seen in the plots to follow. The firstset of plots (Figure 5) show that the attitude tracking error (expressed as a principal angle) and the angular

11 of 14

American Institute of Aeronautics and Astronautics

Page 12: Attitude Tracking Control of a Small Satellite in Low ...mae2.nmsu.edu/~asanyal/Sanyal_res/AIAA-2009-5902-938.pdf · Attitude Tracking Control of a Small Satellite in Low ... The

velocity tracking error go to zero when the control scheme is applied to the system. These plots validatethe effectiveness and robustness of the tracking control scheme which commands the satellite to track thedesired “LVLH hold” attitude and angular velocity even during an instantaneous control input singularity.The control inputs to the actuators, the three magnetic torquer rods and the reaction wheel respectively, canbe seen in the second set of plots (Figure 6). As seen from the plot on the left, one of the magnetic torquerrod inputs, obtained from the control law (24), reaches 20 A·m2, which is 80% of its saturation limit of 25A·m2. When this happens, we set all the actuator inputs constant and equal to their last known value; notethat this keeps the control torque input to the satellite constant. This constant level is maintained until thecontrol input demanded by (24) is below a value of 20 A·m2 for this actuator. Throughout this maneuver,reaction wheel momentum remains less than 4 milli-N·m·s in magnitude, which is way below its momentumsaturation level of 30 milli-N·m·s. The last plot (Figure 7) displays the total required torque magnitude from

Figure 6. Magnetic Torquer moments (left) and reaction wheel momentum (right) required for the satellite de-tumbling

and “LVLH hold” maneuver with a control law singularity

all actuators for this maneuver, as well as the orientation of the geomagnetic field relative to the satelliteduring this maneuver. From this plot of the geomagnetic field, it is clear that the first component of thefield vector expressed in the body frame has a zero crossing a short while after the start of the maneuver.This leads to the singularity in the reaction wheel control law (20). This results in the spike in the requiredinputs for all actuators and the total control torque; however, none of the actuators are staurated becauseof the control logic implemented, which keeps the control inputs to the actuators at a constant level whenone of the magnetic torquer rods reaches 80% of its saturation level. This set of simulation results show thatthe satellite ADCS can be made robust to the instantaneous singularity in the control logic for the reactionwheel. Tracking of the LVLH-hold attitude profile is achieved in less than one-twentieth of an orbit withthis control scheme that averts the singularity in the reaction wheel control law.

VI. Conclusions

In this paper, we presented and applied an almost global attitude and angular velocity tracking controlscheme for a satellite in Low Earth Orbit. This control law is continuous in the attitude and angular velocitystates, and is therefore implementable using actuators that produce continuous torque profiles like reactionwheels and magnetic torquers. This control law is modified and applied to a small 50 kg satellite model in a600 km altitude LEO based on the first planned satellite mission of the Hawaii Space Flight Laboratory. Theattitude actuators in this satellite consist of three magnetic torquers and one small reaction wheel. Analyticaland numerical results show that this controller achieves attitude and angular velocity tracking of “LVLHhold” attitude while recovering from tumbles with large initial tracking errors. The numerical simulationresults also show that the almost global feedback tracking control scheme is robust to atmospheric dragmoments that are not included in the dynamic model used to design the control law. From these simulationresults, we conclude that placement of the reaction wheel is along the roll axis enables the ADCS to maintaincontrollability of the satellite throughout its nominal sun-synchronous orbit. The reaction wheel control inputhas singularity when the local geomagnetic field is normal to the direction of the reaction wheel axis. A

12 of 14

American Institute of Aeronautics and Astronautics

Page 13: Attitude Tracking Control of a Small Satellite in Low ...mae2.nmsu.edu/~asanyal/Sanyal_res/AIAA-2009-5902-938.pdf · Attitude Tracking Control of a Small Satellite in Low ... The

Figure 7. Total control torque magnitude generated by the reaction wheel and magnetic torquer rods (left) and unit

vector components giving the direction of the local geomagnetic field in body frame (right), during the de-tumbling

and “LVLH hold” maneuver with a control law singularity

feedback gain sizing scheme, that varies the attitude and angular velocity feedback gains based on feedback oftrajectory tracking errors, is shown to be effective in carrying out the de-tumbling maneuvers while meetingall actuator saturation limits. When any actuator reaches a certain percentage of its saturation limit dueto the singularity in the reaction wheel control law, all actuator inputs are kept constant till the controllogic reduces the actuator inputs. The simulation results show that the magnetic torquer inputs are morelikely to reach close to their saturation limit than the reaction wheel input. The reaction wheel can providenecessary control effort whenever one of the magnetic torquer rods is aligned with the local geomagneticfield. A reaction wheel momentum dumping scheme could be added to operate periodically to get rid ofexcess momentum building up in the reaction wheel. The future goal of this research would be to apply thistracking control scheme using feedback from attitude and angular velocity sensors that are typically usedin a small satellite. This would combine an attitude determination and estimation scheme with this controlscheme, to create a feedback compensator for the ADCS of a small satellite.

References

1Ahmed, J., Coppola, V. T., and Bernstein, D., “Adaptive Asymptotic Tracking of Spacecraft Attitude Motion withInertia Matrix Identification,” AIAA Journal of Guidance, Control and Dynamics, vol. 21, pp. 684-691, 1998.

2Bhat, S. P., and Bernstein, D. S., “A Topological Obstruction to Continuous Global Stabilization of Rotational Motionand the Unwinding Phenomenon,” Systems and Control Letters, vol. 39, no. 1, pp. 63-70, 2000.

3Bloch, A. M., Baillieul, J., Crouch, P. E., and Marsden, J. E., Nonholonomic Mechanics and Control, Vol. 24 of Seriesin Interdisciplinary Applied Mathematics, Springer-Verlag, New York, 2003.

4Joshi, S. M., Kelkar, A. G., and Wen, J. T., “Robust attitude stabilization of spacecraft using nonlinear quaternionfeedback,” IEEE Transactions on Automatic Control, Vol. 40, No. 10, pp. 1800-1803, 1995.

5Dixon, W. E., Behal, A., Dawson, D. M., and Nagarkatti, S. P., Nonlinear Control of Engineering Systems: A Lyapunov-Based Approach, Birkhauser, Boston, 2003.

6Wen, J. T., and Kreutz-Delgado, K., “The Attitude Control Problem,” IEEE Transactions on Automatic Control, vol.36, no. 10, pp. 1148-1162, 1991.

7Wie, B., Space Vehicle Dynamics and Control, AIAA Education Series, Reston, 1998.8Junkins, J. L., and Schaub, H., Analytical Mechanics of Space Systems, AIAA Education Series, Reston, 1998.9Crassidis, J. L., Vadali, S. R., and Markley, F. L., “Optimal Variable-Sructure Control Tracking of Spacecraft Maneuvers,”

AIAA Journal of Guidance, Control, and Dynamics, Vol. 23, pp. 564-566, 2000.10Cortes, J., “Discontinuous Dynamic Systems,” IEEE Control Systems Magazine, Vol. 28, pp. 3671, June 2008.11Tsiotras, P., “Stabilization and Optimality Results for the Attitude Control Problem,” AIAA Journal of Guidance,

Control, and Dynamics, Vol. 19, pp. 772–779, 1996.12Sharma, R., and Tewari, A., “Optimal Nonlinear Tracking of Spacecraft Attitude Maneuvers,” IEEE Transactions on

Control Systems Technology, Vol. 12, pp. 677–682, 2004.13Koditschek, D. E., “The Application of Total Energy as a Lyapunov function for Mechanical Control Systems,” in

Dynamics and Control of Multibody Systems: Proceedings of the AMS-IMS-SIAM Joint Summer Research Conference atBowdin College, Maine, pp. 131-157, edited by J. E. Marsden, P. S. Krishnaprasad, and J. C. Simo, Vol. 97, AmericanMathematical Society, 1988.

13 of 14

American Institute of Aeronautics and Astronautics

Page 14: Attitude Tracking Control of a Small Satellite in Low ...mae2.nmsu.edu/~asanyal/Sanyal_res/AIAA-2009-5902-938.pdf · Attitude Tracking Control of a Small Satellite in Low ... The

14Chaturvedi, N. A., and McClamroch, N. H., “Almost global attitude stabilization of an orbiting satellite including gravitygradient and control saturation effects,” Proccedings of American Control Conference, pp. 1748-1753, June 2006.

15Sanyal, A. K., and Chaturvedi, N. A., “Almost Global Robust Attitude Tracking Control of Spacecraft in Gravity,”Proceedings of the AIAA Guidance, Navigation, and Control Conference, Honolulu, HI, August 2008, AIAA-2008-6979.

16Sanyal, A. K, Fosbury, A., Chaturvedi, N. A., and Bernstein, D. S., “Inertia-free Spacecraft Attitude Tracking withDisturbance Rejection and Almost Global Stabilization,” AIAA Journal of Guidance, Control, and Dynamics, Vol. 32, pp.1167-1178, 2009.

17Goldstein, H., Classical Mechanics, Second edition, Addison-Wesley, Boston, 1980.18Greenwood, D. T., Classical Dynamics, Second edition, Prentice Hall, Englewood Cliffs, 1987.19Sanyal, A. K., Lee, T., Leok, M., and McClamroch, N. H., “Global Optimal Attitude Estimation using Uncertainty

Ellipsoids,” Systems and Control Letters, vol. 57, no. 3, pp. 236-245, 2008.20Khalil, H. K., Nonlinear Systems, Prentice Hall, Upper Saddle River, NJ, 2002.21Milnor, J., Morse Theory, Princeton University Press, Princteon, NJ, 1963.22Lee, T., Leok, M., and McClamroch, N. H., “A Lie Group Variational Integrator for the Attitude Dynamics of a Rigid

Body with Applications to the 3D Pendulum”, Proc. of IEEE Conf. on Control Applications, pp. 962-967, 2005.23Hussein, I. I., Leok, M., Sanyal, A. K., and Bloch, A. M., “A discrete variational integrator for optimal control problems

on SO(3)”, Proc. of IEEE Conf. on Decision and Control, pp. 6636-6641, 2006.24Picone, J. M., Hedin, A. E., Drob, D. P., and Aikin, A. C., “NRL-MSISE-00 Empirical Model of the Atmosphere:

Statistical Comparisons and Scientific Issues,” J. Geophys. Res., vol. 107, no. A12, pp. 14-68, 2002.25World Magnetic Model, available at: http://www.ngdc.noaa.gov/geomag/WMM/DoDWMM.shtml, based on “NOAA

Technical Report: The US/UK World Magnetic Model for 2005-2010.”

14 of 14

American Institute of Aeronautics and Astronautics