You are on page 1of 12

Previously Published Works

UC Santa Cruz

A University of California author or department has made this article openly available. Thanks to
the Academic Senate’s Open Access Policy, a great many UC-authored scholarly publications
will now be freely available on this site.
Let us know how this access is important for you. We want to hear your story!
http://escholarship.org/reader_feedback.html

Peer Reviewed

Title:
Automatic mass balancing of a spacecraft three-axis simulator: Analysis and experimentation
Journal Issue:
Journal of Guidance, Control, and Dynamics, 37(1)
Author:
Chesi, S
Gong, Q
Pellegrini, V
Cristi, R
Romano Prof, M
Publication Date:
01-01-2014
Series:
UC Santa Cruz Previously Published Works
Permalink:
http://escholarship.org/uc/item/7q4549j9
DOI:
http://dx.doi.org/10.2514/1.60380
Local Identifier:
1002686
Abstract:
Spacecraft three-axis simulators provide frictionless and, ideally, torque-free hardware simulation
platforms that are crucial for validating spacecraft attitude determination and control strategies. To
reduce the gravitational torque, the distance between the simulator center of mass and the center
of rotation needs to be minimized. This work proposes an automatic mass balancing system for
spacecraft simulators, which uses only the three sliding masses during the balancing process,
without need of further actuators. The proposed method is based on an adaptive nonlinear
feedback control that aims to move, in real time, the center of mass toward the spacecraft

eScholarship provides open access, scholarly publishing


services to the University of California and delivers a dynamic
research platform to scholars worldwide.
simulator's centerof rotation. The stability of the feedback system and the convergence of the
estimated unknown parameter (thedistance between the center of mass and the center of rotation)
are analyzed through Lyapunov stability theory. The proposed method is experimentally validated
using the CubeSat Three-Axis Simulator at the Spacecraft Robotics Laboratory of the Naval
Postgraduate School. © 2013 by the authors. Published by the American Institute of Aeronautics
and Astronautics, Inc., with permission.

Copyright Information:
All rights reserved unless otherwise indicated. Contact the author or original publisher for any
necessary permissions. eScholarship is not the copyright owner for deposited works. Learn more
at http://www.escholarship.org/help_copyright.html#reuse

eScholarship provides open access, scholarly publishing


services to the University of California and delivers a dynamic
research platform to scholars worldwide.
JOURNAL OF GUIDANCE, CONTROL, AND DYNAMICS
Vol. 37, No. 1, January–February 2014

Automatic Mass Balancing of a Spacecraft Three-Axis


Simulator: Analysis and Experimentation

Simone Chesi,∗ Qi Gong,† and Veronica Pellegrini‡


University of California, Santa Cruz, Santa Cruz, California 95064
and
Roberto Cristi§ and Marcello Romano¶
Naval Postgraduate School, Monterey, California 93943
DOI: 10.2514/1.60380
Spacecraft three-axis simulators provide frictionless and, ideally, torque-free hardware simulation platforms that
are crucial for validating spacecraft attitude determination and control strategies. To reduce the gravitational torque,
Downloaded by UNIV OF CALIFORNIA - SANTA CRUZ on June 3, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.60380

the distance between the simulator center of mass and the center of rotation needs to be minimized. This work
proposes an automatic mass balancing system for spacecraft simulators, which uses only the three sliding masses
during the balancing process, without need of further actuators. The proposed method is based on an adaptive
nonlinear feedback control that aims to move, in real time, the center of mass toward the spacecraft simulator’s center
of rotation. The stability of the feedback system and the convergence of the estimated unknown parameter (the
distance between the center of mass and the center of rotation) are analyzed through Lyapunov stability theory. The
proposed method is experimentally validated using the CubeSat Three-Axis Simulator at the Spacecraft Robotics
Laboratory of the Naval Postgraduate School.

Nomenclature R = measurement covariance matrix


Er = system rotational energy, J Rbi = rotation matrix between inertial-fixed to body-fixed
Et = system total energy, J coordinate system
G = Jacobian of Y; DYT r = mass balances position vector, m
gb = gravity vector, body-fixed coordinate system, m∕s2 roff = center of rotation to center of mass vector, m
gbx = x component of the gravity vector, body coordinate roff
e = offset estimation error, m
system, m∕s2 roff
x = x component of the center of rotation to center of mass
gby = y component of the gravity vector, body coordinate vector, m
system, m∕s2 rx = center of rotation to mx vector, m
gbz = z component of the gravity vector, body coordinate roff
y = y component of the center of rotation to center of mass
system, m∕s2 vector, m
gi = gravity vector, inertial-fixed coordinate system, m∕s2 ry = center of rotation to my vector, m
H = measurement distribution matrix roff
z = z component of the center of rotation to center of mass
J = spacecraft simulator inertia matrix, kg · m2 vector, m
J0 = spacecraft simulator inertia matrix before mass rz = center of rotation to mz vector, m
balancing procedure Ts = sample time, s
kp = arbitrary positive scalar t = time, s
mp = sliding masses mass, kg V = Lyapunov function
ms∕c = spacecraft simulator mass, kg v = measurement noise
mx = sliding mass in the xb direction w = process noise
my = sliding mass in the yb direction X = state vector for q; ω; Θ
~ T
mz = sliding mass in the zb direction Xi = x coordinate, inertial coordinate system, m
Pg = system potential energy, J x = state vector for ωx ; ωy ; ωz ; roff
z 
T

Pp = projection factor xb = x coordinate, body coordinate system, m


P0 = error covariance matrix Y = state vector for ω
Q = process covariance matrix Yi = y coordinate, inertial coordinate system, m
q = quaternion vector yb = y coordinate, body coordinate system, m
Zi = z coordinate, inertial coordinate system, m
zb = z coordinate, body coordinate system, m
Received 28 September 2012; revision received 21 February 2013; Γ = spacecraft simulator angular momentum, N · m · s
accepted for publication 11 September 2013; published online 23 October Θ = offset vector, m
2013. Copyright © 2013 by the authors. Published by the American Institute Θ^ = offset estimation, m
of Aeronautics and Astronautics, Inc., with permission. Copies of this paper Θ~ = offset estimation error, m
may be made for personal or internal use, on condition that the copier pay the
σ = standard deviation
$10.00 per-copy fee to the Copyright Clearance Center, Inc., 222 Rosewood
Drive, Danvers, MA 01923; include the code 1533-3884/13 and $10.00 in τr = theoretical control torque, N
correspondence with the CCC. ω = angular velocity vector, rad∕s
*Ph.D. Candidate, Applied Mathematics and Statistics Department; ωg = angular velocity vector parallel to the gravity field, rad∕s
schesi@soe.ucsc.edu. Member AIAA. ωp = angular velocity vector perpendicular to the gravity

Associate Professor, Applied Mathematics and Statistics Department; field, rad∕s
qigong@soe.ucsc.edu. Member AIAA.

Ph.D. Candidate, Applied Mathematics and Statistics Department;
vpellegr@soe.ucsc.edu.
§ I. Introduction
Professor, Electrical Engineering Department; rcristi@nps.edu.

Associate Professor, Mechanical and Aerospace Engineering Department;
mromano@nps.edu. Associate Fellow AIAA. E XPERIMENTAL test beds to simulate spacecraft dynamics,
navigation, and control can greatly reduce the costs and risks
197
198 CHESI ET AL.

related to the design of attitude control and determination systems


for small spacecraft [1]. A three-axis simulator consists of a
hollow hemispherical air bearing, floating over a correspondent
hemispherical cup through which compressed air is flowing. The air-
bearing hemisphere contains sensors, actuators, computers, and
power storage/conditioning to fully model a spacecraft attitude
determination and control system. The simulator enables full three-
degree-of-freedom rotational motion for attitude dynamics and
control simulations [2]. To simulate a frictionless and torque-free
environment for rigorous ground testing of spacecraft systems,
disturbance torques must be compensated. The major disturbance
torque that spoils the rigorous reproduction of a torque-free space
environment is due to gravity [3]. To minimize this disturbance
torque, the distance between the center of mass and the center of
rotation needs to be minimized. A standard procedure for manually
balancing a spacecraft simulator is described in [4], where, by
inspecting the spacecraft simulator pendulum motion, the balance
masses are moved to increase the period of the pendulum motion. Fig. 1 Automatic mass balancing system.
Downloaded by UNIV OF CALIFORNIA - SANTA CRUZ on June 3, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.60380

This method requires multiple trials and is limited in real applications


because of the rotational travel constraints of the spherical air
bearing. As shown in [5], by analyzing different points of simple to implement, because only three sliding masses moved by
equilibrium, it is possible to determinate the relationship between the three motors are needed.
spacecraft center of mass and the known positions of the balance The proposed automatic balancing idea requires a control
masses. However, this method requires a considerable amount of algorithm that automatically adjusts the location of the sliding masses
time and does not guarantee the correct estimation of the center of to eliminate the unknown offset between the center of rotation and
mass. Manual balancing has also been implemented in [6], where the center of mass. There are some challenges that need to be successfully
disturbance gravitational torque has been reduced to 0.01 Nm. addressed. First, the torque that can be generated by the actuators (the
An automatic mass balancing system is highly desirable for balance masses) is physically confined in the direction perpendicular
spacecraft simulators, particularly during the design process, because to the gravity field. It implies that the offset along the gravity field is
the simulators may be subjected to modifications or upgrades. In fact, very difficult to be compensated using sliding masses. For example,
even a slight modification or relocation of the system components suppose the spacecraft is balanced in the two directions perpendicular
requires a rerun of the manual balancing procedure, which is a time- to the gravity field, and the spacecraft is spinning along the direction
consuming process. Automatic mass balancing systems, on the other of the gravity field. In such situation, the location of the sliding mass
hand, can significantly reduce the balancing time from hours to on the direction of the gravity field does not affect the motion.
possibly minutes. To overcome the limitations of manual balancing Therefore, the control system has very limited feedback information
system, a number of spacecraft simulators have been equipped with to use. To overcome this challenge, a two-step design has been
an automatic balancing system [7–11]. An automatic mass balancing chosen. In the first step, the offset in the transversal direction is
system was proposed by Kim and Agrawal [7], where external compensated using an adaptive nonlinear control law. Once the
control moment gyros were used to track the angular momentum and offset in the two directions perpendicular to the gravity field are
determine the positions of three proof masses. Experiments on an compensated, the offset in the last direction is estimated using a
automatic mass balancing technique were conducted in [10], where Kalman filter and then compensated according to the estimated
the system moves the balance masses to compensate the imbalance offset.
derived from the difference in total impulse exerted by the actuators In the first step of the automatic balancing procedure (i.e.,
during limit cycle. There are also automatic mass balancing systems compensation of offset in the directions perpendicular to the gravity
that have been designed based on least-squares estimation of the field), due to the presence of unknown parameters (the components of
center of mass [12–14]. The limitation of such least-squares-based the distance vector between the spacecraft simulator’s center of mass
techniques is that the estimation and compensation process must be and the center of rotation), a nonlinear adaptive control is adopted.
repeated several times before accurate balancing is achieved. Conceptually speaking, the proposed control algorithm is based on
This paper presents an automatic mass balancing technique, which the conservation of the angular momentum. If the spacecraft
uses exclusively three balance masses without any other actuators. simulator is perfectly balanced, the angular momentum is conserved
The proposed automatic balancing system is based on adaptive because there is no acting external torque. Conversely, if an offset is
nonlinear feedback control that significantly and automatically present between the center of mass and the center of rotation, the
reduces the gravitational disturbance torque by a precise collocation angular momentum is not conserved due to the gravitational torque.
of the center of mass on the center of rotation. The automatic mass This information can be used by the control system to relocate the
balancing system is composed of three balance masses that are moved positions of the three sliding masses until the derivative of the angular
by electric motors along the three orthogonal directions of the body momentum is zero. The proposed nonlinear adaptive control law is
coordinate system. By assumption, the body coordinate system is proven to be stable in the sense of Lyapunov. Furthermore, by using
aligned with the spacecraft simulator’s principal inertia axes with the LaSalle’s invariance principle, it has been proven that the offsets
origin located in the spacecraft simulator’s center of rotation (CR) as along the two directions perpendicular to the gravity field can be
shown in Fig. 1. The vector roff   roff
x roff
y roff T
z  represents the completely compensated. In the second step, the vertical offset is
unknown offset between the CR and the center of mass (CM). The estimated by an unscented Kalman filter and then compensated.
goal of the automatic mass balancing system is to eliminate the The proposed automatic mass balancing system is tested on a
unknown offset by moving the three balance masses mx , my , and mz newly developed spacecraft simulator at the Spacecraft Robotics
in an appropriate manner. Laboratory at the Naval Postgraduate School [15]. The purpose of
With respect to other automatic balancing systems, the main this apparatus is to simulate the attitude dynamics of a class of small
advantage of the proposed system is that it does not require any other satellites: CubeSats. The simulation and experimental results are in
actuators besides balance masses to balance the spacecraft simulator. strong agreement with the theoretical analysis and demonstrate a
This aspect is particularly important when dealing with small significant reduction of the disturbance torque.
spacecraft simulators, because actuators, such as reaction wheels or The paper is organized as follows. The first step of the balancing
control moment gyros, require a consistent amount of volume and procedure is described in Sec. II. This step uses a nonlinear adaptive
weight. Moreover, the proposed automatic balancing system is control law to compensate the transversal imbalance. The vertical
CHESI ET AL. 199

imbalance compensation based on estimation techniques is to the CR. Furthermore, due to the small entity of the masses and the
considered in Sec. III. Section IV presents the simulation results, low speed of their motion, it has been assumed that the effect of the
and the experimental results on the spacecraft simulator are presented motion of the balancing masses on the angular momentum can be
in Sec. V. Finally, Sec. VI states the conclusions of this work. neglected. In Eq. (4), ri × is a cross-product matrix defined as
2 3 23
II. Transversal Imbalance Compensation 0 −r3 r2 r1
r×  4 r3 0 −r1 5 with r  4 r2 5 (5)
A. System Dynamic Model
−r2 r1 0 r3
The spacecraft simulator kinematics and dynamics can be modeled
using two coordinate systems as shown in Fig. 2: the inertial
coordinate systems Xi , Y i , Zi and a coordinate system fixed to the To balance the spacecraft simulator, the Earth gravitational field gb in
simulator body xb, yb , zb . Both coordinate systems have their origins the body coordinate system must be computed using the simulator
in the center of rotation. The orientation of the body-fixed coordinate attitude through the direction cosine matrix
system with respect to the inertial coordinate system defines the
attitude of the spacecraft simulator. gb  Rbi gi (6)
The spacecraft three-axis simulator can rotate freely, but cannot
translate, because the CR is a fixed point in the inertial coordinate where gi   0 0 −9.81 T m∕s2 is the gravitational field in the
system. Each balance mass is assumed to be an ideal mass particle inertial coordinate system, and Rbi is the rotation matrix defined as
Downloaded by UNIV OF CALIFORNIA - SANTA CRUZ on June 3, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.60380

that can be translated along its corresponding axis. The rotational 2 3


kinematics using a quaternion can be describe by 1 − 2q22  q23  2q1 q2 − q4 q3  2q1 q3  q4 q2 
2 3 2 32 3 Rbi  4 2q1 q2  q4 q3  1 − 2q21  q23  2q2 q3 − q4 q1  5
q_ 1 0 ωz −ωy ωx q1 2q1 q3 − q4 q2  2q1 q3  q4 q1  1 − 2q21  q22 
6 q_ 2 7 1 6 −ωz 0 ωx ωy 7 6 7
6 7 6 76 q2 7 (1) (7)
4 q_ 3 5 2 4 ωy −ωx 0 ωz 54 q3 5
q_ 4 −ωx −ωy −ωz 0 q4 The purpose of this paper is to use the sliding masses to relocate
the CM of the spacecraft simulator toward the CR to eliminate
where ω   ωx ωy ωz represents the absolute angular velocity
T the gravitational torque. The proposed method assumes that the
of the simulator along the body coordinate system, and q  simulator inertia matrix is known and constant. In the next sections, a
 q1 q2 q3 q4 T are the quaternions. two-step design is presented. In the first step, the two transversal
It is assumed that the system subjects to no external control torque directions xb and yb are balanced by using an adaptive nonlinear
except for the gravity torque due to the system imbalance. Assuming feedback control; in the second step, the imbalance in the remaining
the balance masses are held in their original positions, the rotational direction is estimated by using an unscented Kalman filter, and then
dynamics of the simulator with respect to the center of rotation is the offset is compensated. The main reason for such a two-step design
given by Euler’s equation [16] is that the torque generated by the moving masses is always
constrained in a plane orthogonal to the gravitational field, which
Γ_  ω × Γ  roff × ms∕c gb (2) greatly limits the control space and generates difficulties in feedback
control design.
where ms∕c is the simulator mass including the three proof masses,
and Γ is the angular momentum, which can be written as B. Adaptive Nonlinear Feedback Control Law
The proposed adaptive control technique is based on the
Γ  Jω (3)
conservation of the angular momentum. If the gravitational torque is
zero, the angular momentum is conserved. However, in the presence
with J as the simulator inertia matrix. By assumption, the body
coordinate system is parallel to the principal inertia axes of the of an offset between CM and CR, and therefore of a gravitation
simulator, hence the inertia matrix is diagonal. Additionally, it has torque, the angular momentum is not conserved. The main concept of
been assumed that the three balance masses are aligned with the axes the proposed imbalance compensation method is to impose that the
of principal inertia and that they have the same mass mp . The inertia derivative of the angular momentum become zero by displacing, in
matrix can then be written as real time, the three balance masses.
The mechanical control generated by moving the balance masses
X can be expressed as
J  J0 − mp ri ×ri × for i  x; y; z (4)
i X
τ r  mp ri × gb for i  x; y; z (8)
where J0 is the inertia matrix before the balancing procedure, and the i

vector ri represents the position of the ith balance mass with respect
where ri , the ith balance mass’s position with respect to the CR, is the
control input to be designed.** By considering Eq. (2) and taking into
consideration the effect of τ r , the system dynamics can be written as

Γ_  ω × Jω  roff × ms∕c gb  τ r (9)

In this formulation, the torque τ r will be treated as the dummy control


to be designed using nonlinear adaptive control. Then, the real
control ri will be designed to generate the desired control torque.
However, from Eq. (8), clearly the torque generated by moving the

**Notably, the effect of the displacement of the three equal masses, sliding
each in one direction, on the generation of the gravitational torque, is the same
that it would be obtained by using a single mass displaced in three dimensions
by the sum of the displacement vectors of the three masses. However, using
three masses has the advantage of leaving diagonal the matrix of principal
Fig. 2 Inertial and body coordinate systems. inertia with respect to the center of rotation.
200 CHESI ET AL.

masses is constrained in the plane orthogonal to gb . This infers that, in V_  ωT ΦΘ


^ − ωT ΦΘ
^ − kp ωT ωp
the design of the control torque τ r , it needs to contain no component
along the vector gb.  −kp ωT ωp
To design the theoretical torque τ r , let Θ  roff , where the  −kp ωg  ωp T ωp (21)
parameter roff represents the unknown and time-invariant initial
offset between the center of mass and the center of rotation. Then, Since ωTg ωp  0, this yields
Θ × ms∕c gb  ΦΘ (10)
V_  −kp kωp k2 (22)
with From Eq. (22), it is possible to see that the derivative of the Lyapunov
function is negative semidefinite, hence the closed-loop control
Φq  −ms∕c gb × (11) system is Lyapunov stable. Moreover, by the LaSalle invariance
principle [17], the system will converge to the largest invariant set Ω
Let Θt
^ be the estimation of the unknown vector parameter Θ, and contained in
define the estimation error Θt
^ as
_ ≡ 0g  fX∶Vq;
fVt ~ ≡ 0g  fωp t ≡ 0g
_ ω; Θ (23)
Θt
~  Θ − Θt
^ (12)
Downloaded by UNIV OF CALIFORNIA - SANTA CRUZ on June 3, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.60380

in other words, limt→∞ ωp t  0.


The dynamics of Θt^ will be determined later through Lyapunov
analysis. C. Control Torque Generation
Define the state vector for the overall problem as X  q; ω; Θ
~ T
The designed theoretical control torque τ r must be physically
The angular velocity vector ω can be decomposed into two
created by moving the balance masses. The torque generated by the
components, one ωg along the vector gb and the other ωp, orthogonal sliding masses is constrained to a direction normal to both the masses’
to gb , that is, position vector and the gravitational field direction. If τ r is designed
according to Eq. (20), this property can be guaranteed. In fact,
ω  ωg  ωp ; ωTg ωp  0; gb T ωp  0 (13) multiplying gb T to both sides of Eq. (20), and using Eqs. (11) and
(13), leads to
Let us also define the projection operator
  gb T τ r  −gb T ΦΘ
^ − kp gb T ωp  0 (24)
b b T
P p q  I − gkggb k2 (14)
To design the real control r, which is the balance masses’ position
vector, the designed torque τ r needs to be mapped to r. For this
so that purpose, substituting τ r in Eq. (8) gives

ωp  P p qω (15) τ r  mp −gb t × r (25)

where ωp represents the angular velocity perpendicular to the Earth’s Because the matrix −gb t× is always singular, r cannot be directly
gravitational field. found by inverting it. However, due to property (24), the designed
To design the state feedback control and the adaptive law for Θ,
^ the control torque τ r is guaranteed to be in the range of −gb t× for
following positive definite Lyapunov function is used: all t, so that Eq. (25) always has a solution for r. Indeed, a solution is
given by
Vq; ω; Θ
~  1 ωT Jω  1 Θ
2 2
~ TΘ
~  1 qT q
2
(16)
r  kggb k×τ2 mr
b

p
(26)
The time derivative of the Lyapunov function along the motion of the
system leads to To see this, substituting Eq. (26) into Eq. (25), and using the condition
gb ⊥ τ r , ∀ t from Eq. (24), it can be found that
_  ωT Jω
_ Θ _~
~ TΘ
Vt  qT q_    
g b × gb × τ r gb · τr gb − gb · gb τ r
 ω −ω × Jω  ΦΘ  τ r   Θ
T _~
~ TΘ −mp  −mp  τr
kgb k2 mp kgb k2 mp
_~
 ωT ΦΘ  ωT τ r  Θ
~ TΘ (17) (27)

where Eqs. (9) and (11) have been used, together with the hypothesis Therefore, the control applied to the mass balancing system is given
by Eq. (26) with τ r be designed according to Eq. (20).
of constant J. From Eqs. (12) and (17) became

_~ D. Convergence of the Estimated Parameters


V_  ωT ΦΘ
^ Θ
~ T ΦT ω  Θ  ωT τ r (18)
By the LaSalle invariance principle [17], the system will converge
to the largest invariant set Ω contained in Eq. (23). On the set
Let the adaptive law for the estimated parameter be fωp t ≡ 0g, the designed parameters estimation law in Eq. (19)
leads to
_^
Θ  ΦT ω (19)
_^
Θ  ΦT ω  ΦT ωp  ΦT ωg  ΦT ωg  0 (28)
and the control torque be
because ωg is always orthogonal to all columns of Φ, which
τ r  −ΦΘ
^ − kp ωp (20) represents the cross-product matrix form of the gravity field.
On the set fωp t ≡ 0g, the spacecraft remains to spin with an
where the parameter kp is an arbitrary positive scalar. Substituting angular velocity aligned to the gravity field. This condition is shown
Eqs. (19) and (20) into Eq. (18), obtaining in Fig. 3, where it has been assumed that the spacecraft remains to
CHESI ET AL. 201

J−1
xx · gz · Θy  0
b ~ − J−1
yy · gz · Θx  0 0 · Θz  0
b ~ ~ (37)

Since J−1 −1
xx , J yy , gz ≠ 0, Eq. (37) implies Θx  Θy  0. Thus, rx ,
b ~ ~ off

roff
y have been correctly estimated and compensated. In fact, by
considering Eqs. (20) and (23), the right-hand side of Eq. (9) is equal
to zero and this implies that no gravitational torque is acting on the
spacecraft (i.e., the distance between CM and CR has been
compensated at least in two directions). But nothing can be claimed
about the third component of Θ. ~ This component will be estimated
using a Kalman filter as demonstrated in the following section.

III. Vertical Imbalance Compensation


Fig. 3 Spacecraft simulator zb axis aligned with gb .
Once the offsets along roff off
x and ry directions are balanced, the
b
offset in the remaining z direction needs to be compensated.
rotate around the zb axis. Under this condition, it is possible to prove Spacecraft attitude and parameter estimation has been studied using
the convergence of the parameter. Indeed, the closed-loop dynamics the extended Kalman filter (EKF) and unscented Kalman filter (UKF)
Downloaded by UNIV OF CALIFORNIA - SANTA CRUZ on June 3, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.60380

can be obtained by replacing the control actions into Eq. (2) so that [18–21]. For the dynamic system described in Eq. (2), either UKF or
EKF can be used, and the performance is expected to be similar. In
this study, the UKF is used to estimate the offset roff
z by a standard
_  J−1 −ω × Jω  ΦΘ
ω ~ − kp ωp  (29) approach of augmenting this unknown parameter into the state
space [22].
The projection factor P p in Eq. (14), can be written as To write the system in state-space form, let us define the state
2 3 2 b 2 3 vector of the continuous system as x  ωx ; ωy ; ωz ; roff T
z  . Here, the
1 0 0 gx  gbx gby gbx gbz off
unknown parameter rz is augmented into the state with the
Pp q  4 0 1 0 5 − 4 gby gbx gby 2 gby gbz 5 kg1b k (30) corresponding dynamics to be r_off
z  0. The disturbance torque due
0 0 1 gbz gbx gbz gby gbz 2 to the imbalance along the zb axis is

Because on the set Ω, ωp t ≡ 0, the spacecraft simulator is rotating roff × ms∕c gb   −ms∕c gby roff b off
z ;s∕c gx rz ; 0 
T
(38)
around the gravity field along the zb axis so that the components gbx ,
gby of the projection factor go to zero. As a consequence, Therefore, the spacecraft simulator dynamics in the state-space form
is given by
2 3
1 0 0 x_  fx  dt (39)
Pp q  4 0 1 0 5 (31) where
0 0 0
2 0 2 31 3
−ms∕c · gby · roff
z
Premultiplying Eq. (29) by the projection factor P p leads to 6 J−1 @−ω × Jω  4 m · gb · roff 5A 7
f 6
4
s∕c x z 7
5 (40)
0
dP p ω
Pp ω
_  _ p  Pp J −1 −ω × Jω  ΦΘ
ω ~ − K p ωp  (32) 0
dt
and dt represents the process noise. Let the measurement be the
where the time derivative of the projection factor is equal to zero only angular velocity given by the inertial measurement unit (IMU). Then
on the set fωp t ≡ 0g. Since ω  ωg  ωp on fωp ≡ 0g, Eq. (32) the output equation is given by y  Hx  vt, where
becomes
2 3
1 0 0 0
0  Pp J−1 −ωg × Jωg  ΦΘ
~ (33) H  40 1 0 05 (41)
0 0 1 0
Because J is diagonal and ωg has only one component along the zb
axis, it can be shown that ωg × Jωg  0. Therefore, Eq. (33) and vt represents the measurement noise.
becomes To correctly estimate the offset in the last direction using the UKF,
the observability of the unknown parameter needs to be analyzed. In
0  Pp J−1 ΦΘ
~ (34) this paper, the definition of observability for the nonlinear system
introduced in [23] has been adopted. Consider a general nonlinear
where the matrix Φ can be written, from Eq. (11), as system
2 3 ξ_  ft; ξ; u; ξ ∈ Rn ; u ∈ Rm Y  ht; ξ; u (42)
0 gbz 0
Φ  ms∕c 4 −gbz 0 05 (35) where ξ is the state, ut is a continuous control input, and Y
0 0 0 represents the variable that can be directly measured by sensors.
Suppose z  zt; ξ; u is a variable to be estimated. Let U represent
And so, Eq. (34) can be written as an open and connected set in the time-state-control space
R × Rn × Rm .
2 32 −1 3 32 3 Definition 1 (Observability) [23]: The function z  zt; ξt; ut
1 0 0 Jxx 0 0  0 gbz 0 Θ~ x
is said to be observable in U if for any two trajectories t; ξi t; ui t,
0  ms∕c 4 0 1 0 54 0 J−1
yy 0 5 −gbz 0 0 54 Θ~ y 5
i  1; 2 in U defined on a same interval t0 ; t1 , the equality
0 0 0 0 0 J−1
zz 0 0 0 Θ~z
(36) hξ1 t; u1 t  hξ2 t; u2 t; a:e: in t0 ; t1  (43)

which yields implies


202 CHESI ET AL.

zt; ξ1 t; u1 t  zt; ξ2 t; u2 t (44) graduate School. The spacecraft simulator inertia matrix is
J  diag0.0226; 0.0257; 0.0266 kg · m2 . The total mass is
almost everywhere in t0 ; t1 . Suppose for any trajectory t; ξt; ut ms∕c  4.2 kg with the balance masses mx  my  mz  0.3 kg.
in U there always exists an open set U1 ⊂ U so that t; ξt; ut is An offset is simulated with values roff  1 − 0.9 − 1.4T · 10−3 m,
contained in U1 and zt; ξ; u is observable in U1 . Then, z  the initial angular velocity ω   0.0888 0.8229 1.3611 T rad∕s,
zt; ξ; u is said to be locally observable in U. and quaternion q   0 0 0 1 T . In the simulation, the sliding
Lemma 1 [23]: Consider a system without control masses are used as actuators. The feedback control law in Eq. (20) is
generated using the simulated data of angular velocity, gravity field,
ξ_  ft; ξ; ξ ∈ Rn Y  ht; ξ (45) and quaternions. Furthermore, the simulation assumes that the data
related to the gravity field, angular velocity, and quaternions have the
Let U ⊂ R × Rn be an open set. Consider same sample time T s  0.05 s. The maximum translation distance
for each sliding mass is 28 mm providing a total offset com-
V  Y T ; DY T ; : : : : : : ; Dl−1 Y T T (46) pensation capability of 2 mm in each direction. The uncertainty in
the sliding masses’ positions has been modeled using Gaussian white
for some l > 0, where D is the differentiation operator. If noise distribution with standard deviation σ  0.5 · 10−6 m. More
  details on the CubeTAS simulator are presented in Sec. V.
∂V In the first step of the proposed automatic balancing procedure, the
rank n
∂ξ adaptive control designed in Sec. II is applied to eliminate the
Downloaded by UNIV OF CALIFORNIA - SANTA CRUZ on June 3, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.60380

transversal imbalance. The simulation results are demonstrated in


for t; ξ ⊂ U, then z  zt; ξ is locally observable in U. Figs. 4–6. Figure 4 shows the balance masses’ positions during the
The aforementioned observability definition and Lemma are balancing procedure. As a physical constraint, the sliding masses’
applied to the spacecraft simulator dynamics in Eq. (40) with displacements need to be less than the maximum travel distance of
measurement ht  ωx ; ωy ; ωz . The variable to be estimated is 28 mm for all time. For this reason, the feedback gain kp in Eq. (20) is
zt; ξ  roff
z . By denoting the differentiation operator D, then

Y  ω 0.02
2 0 2 31 3 0.01
−ms∕c · · gby roff
z
6 −1 B 6 7C 7 0
_ 6
DY  ω J B−ω × Jω  6 ms∕c · gbx · roff 7C 7 (47)
4 @ 4 z 5A 5 −0.01
−0.02
0 0 50 100 150 200 250 300 350 400

According to Lemma 1, if the Jacobian of Y; DYT , that is, 0.02


  0.01
∂ Y
G (48) 0
∂X DY −0.01
−0.02
has full rank, then local observability can be guaranteed. The 0 50 100 150 200 250 300 350 400
Jacobian matrix has the following triangular structure:
  0.02
I 3×3 03×1 0.01
G (49)
G21 G22 0
−0.01
where I3×3 is the (3 × 3) identity matrix, G21 is a (3 × 3) matrix, and −0.02
G22 is a 3 × 1 vector given by 0 50 100 150 200 250 300 350 400

2 3 Fig. 4 Simulation results: balance masses’ positions.


m gby
− s∕c
∂ω_ 6 Jxx 7
G22  off  6 m gbx 7
4 − s∕c 5 (50)
∂rz Jyy −3
x 10
0 2
1
0
Hence, the Jacobian matrix G has full column rank (i.e., the system is −1
locally observable), if gby ; gbx  ≠ 0; 0. This condition can be satisfied −2
if the simulator is tumbling under the gravity torque. For this reason, −3
0 50 100 150 200 250 300 350 400
to estimate the unknown parameter roff z , initially the simulator has
been put in a tumbling motion by giving it an angular velocity with x 10
−3

components in all three directions. After roff 2


z has been estimated, the 1
balance mass mz can be moved along the zb axis to compensate the 0
offset in that direction, thus concluding the automatic balancing in all −1
three directions. −2
−3
0 50 100 150 200 250 300 350 400

IV. Simulation Results x 10


−3
2
The simulation results shown here validate the effectiveness of the 1
balancing procedure proposed earlier. All the simulations are 0
developed using MATLAB/Simulink 7.0 using the Runge–Kutta −1
fourth-order integrator with fixed step T s  0.05 s. The physical −2
−3
parameters used for this simulation are based on the real parameters 0 50 100 150 200 250 300 350 400
of the CubeSat three-axes simulator (CubeTAS) developed
at the Spacecraft Robotics Laboratory at the Naval Post- Fig. 5 Simulation results: convergence of the estimated offsets.
CHESI ET AL. 203

2.4 0.03
1.8
1.2 0.02
0.6
0 0.01
−0.6
0 50 100 150 200 250 300 350 400 0

0.01
2.4 0 5 10 15 20
1.8
1.2
−5
0.6 x 10
0 2
−0.6
0 50 100 150 200 250 300 350 400 1.5

1
2.4
1.8 0.5
1.2
0.6 0
0 0 5 10 15 20
Downloaded by UNIV OF CALIFORNIA - SANTA CRUZ on June 3, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.60380

−0.6
0 50 100 150 200 250 300 350 400
Fig. 7 Simulation results: Top graph shows the estimation of roff
z and
the real value. Lower graph shows the estimation error roff
e using UKF.
Fig. 6 Simulation results: magnitude of ω, ωp and ωg during
transversal balancing procedure.
estimated, and then it can be compensated to drastically reduce the
gravitational torque acting on the spacecraft simulator.
tuned in a way that the sliding mass travel distance is always less than
To experimentally verify the effectiveness of the automatic
its maximum value, as shown in Fig. 4.
balancing procedure proposed here, the following method will be
As shown in Fig. 5, during the mass balancing procedure, Θ~ x and
implemented on the CubeTAS. After the balancing procedure, an
Θ~ y converge to zero, implying that the offsets in xb and yb directions
arbitrary angular velocity will be applied to the system. The total
are balanced. Θ~ z does not converge to zero, because the offset in the
energy, which is the sum of the kinetic energy and the potential
third direction roff
z cannot be balanced with the proposed adaptive energy, is conserved, that is,
control. These simulation results agree with our analysis presented in
preceding sections. Figure 6 shows the behavior of the angular
Etot  Er t  Pg t  constant (55)
velocity during the first step of the balancing procedure. The angular
velocity perpendicular to the gravity field ωp tends to zero and
ω → ωg . Thus, the spacecraft remains to spin along the gravity vector where Pg t represents the potential energy and Er t 
2 ω tJωt represents the kinetic rotational energy. If the system
1 T
as proven in Sec. II using Lyapunov stability analysis.
Once the first step is completed, the offset in the directions is balanced, then Pg t ≅ 0. Therefore, the rotational kinetic energy
perpendicular to the gravity field has been compensated. The offset in is conserved during the motion, that is, Er t  Er 0  constant.
the last direction roff Thus the variation of the rotational kinetic energy in time serves as a
z can be estimated in the second step. At the
beginning of the second step, an angular velocity with components in measure of the effectiveness of the mass balancing procedure.
all directions is applied to the spacecraft simulator. This guarantees Because the rotational kinetic energy can be evaluated by the
the observability of the unknown parameter roff gyroscope measurement, this quantity provides an easy way to
z as explained in
Sec. III. During the rotational motion, a UKF is adopted to estimated demonstrate the accuracy of the balance procedure experimentally.
roff Figure 8 shows the variations of the kinetic energy before and after
z . The associated measurement and process covariance matrices,
considering the hardware setup, have been chosen as the balancing procedure. When the system is unbalanced, the
variation in amplitude of the kinetic energy will be greater than the
amplitude variation with respect to the balanced system. As shown in
R  EvvT   diag4.87 × 10−5 2 ; 4.87 × 10−5 2 ;
51 Fig. 8, the kinetic energy of the unbalanced system has a standard
4.87 × 10−5 2 rad2 ∕s2 
deviation σ unbalance  3.847 × 10−3 J, where the balanced system
has a standard deviation σ balance  4.5331 × 10−6 J; thus this value
Q  EddT   diag10−10 ; 10−10 ; 10−10 ; 1 × 10−4 2 
(52)
×rad2 ∕s4 ; rad2 ∕s4 ; rad2 ∕s4 ; m2 ∕s2  0.02

The initial error covariance matrix P0 , reflecting the uncertainties in 0.018

the initial estimated state variables, is set to be a diagonal matrix as


0.016

P0  diag0.12 ; 0.12 ; 0.12 ; 1.5 × 10−3 2 


(53)
rad2 ∕s2 ; rad2 ∕s2 ; rad2 ∕s2 ; m2  0.014

0.012
The simulation results using the preceding gain matrices are shown in
Fig. 7, where the error between roff z and its estimation has been
0.01
defined as

e  krz − r^z k
0.008
roff off off (54)
0.006
It can be seen that the filter converges to the correct value of roff
z in less
−6
than 20 s. Indeed, roff
e decreased to less than 10 m after the filter
0.004
converges. The small error in the estimation shown in Fig. 7 depends 0 5 10 15 20 25 30 35 40
on the UKF sample time T s  0.05 s. In fact, simulations show that a
reduction of the sample time leads to a reduction of roff e . Thus, it can Fig. 8 Simulation results: kinetic energy before Ebr and after Ear the
be claimed that the offset in the last direction has been correctly balancing procedure.
204 CHESI ET AL.
Downloaded by UNIV OF CALIFORNIA - SANTA CRUZ on June 3, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.60380

Fig. 9 CubeSat three-axis simulator at the Spacecraft Robotics Laboratory of the Naval Postgraduate School (Lab Director, Marcello Romano).

has been reduced by approximately 99.9%. The value of the standard of all the components were determined in three-dimensional (3-D)
deviation after balancing procedure is different from zero because the CAD software NX 7 by Siemens. The property of each component
uncertainties in the sliding masses’ positions have been modeled as was derived by the CAD file provided by the component’s
Gaussian white noise distribution. By adding noise into the sliding manufacturer. The CubTAS 3-D CAD model is shown in Fig. 10. The
masses’ positions, the system is subjected to a torque that prevents the upper part is mainly for three reaction wheels and a battery, whereas
kinetic energy from being a constant. However, the sensible the lower part is for PC-104 boards, power management boards, and
reduction in the kinetic energy standard deviation is an indication of balancing masses with relative linear actuators. The IMU provides
an effective balance of the system. information regarding angular velocity and magnetic and gravity
fields. A magnetometer, sun sensor, three magnetic coils, and three
reaction wheels are used for attitude estimation and control. During
V. Experimental Results the automatic mass balancing experiment, only the data relative to
A. Experimental Apparatus
angular velocity and gravity field are used.
CubeTAS is a CubeSat-scale three-axis hardware simulator
B. Automatic Mass Balancing Experiments
designed to recreate an ideally torque-free platform for testing
nanosatellites’ attitude determination and control algorithms in a lab In the first step of the balancing procedure, the offset along the xb
environment. The experimental apparatus shown in Fig. 9 has been and yb directions will be compensated. At the beginning of the first
developed at the Spacecraft Robotics Laboratory of the Naval experiment, an initial angular velocity of ω   0.0888 0.8229
Postgraduate School. 1.3611T rad∕s is applied to the spacecraft simulator. As the
The CubeTAS simulator contains an automatic balancing system, designed adaptive nonlinear control takes action, the IMU registers
consisting of three stainless steel masses driven by three linear the gravity field components and angular velocity values. Shown in
stepper motors along their axis, respectively. The sliding masses are Fig. 11 are the gravity field vector components during the rotational
custom made to encapsulate the stepper motors. The whole system is motion. It can be seen that, after t  200 s, the only component of the
supported by a spherical air bearing that allows it to rotate 50 deg gravity field that is different from zero is the one aligned to the zb axis.
about two axes in the horizontal plane and 360 deg along the vertical This is in agreement with our theoretical analysis. Figure 12 shows
axis [24]. A Helmholtz cage has been constructed around the the magnitude of the components of the angular velocity. Solid lines
spacecraft simulator to allow the simulation of the orbital magnetic are the experimental results and dashed lines are simulation results.
field. Attached to the Helmholtz cage, a high-accuracy motion As shown in Fig. 12, the experimental data relative to the magnitude
capture camera system measures the attitude. The main design goal of the angular velocity ωp tend to zero, so that, at the end of the first
for CubeTAS is to have a shape suitable to be fit in a sphere as shown step of the balancing procedure, the spacecraft simulator remains to
in Fig. 10. The mechanical design was made such that the offset spin only along the zb axis. This behavior was expected from the
between the center of mass and the center of rotation is minimum. To simulation results of the proposed balancing method as shown by the
meet these goals, the physical shape was designed and the locations dashed lines.

Fig. 10 Overview of the CubeTAS body.


CHESI ET AL. 205

4
0
so that each diagonal element reflects the square of each experimental
−4 initialization error. By analyzing the gyroscope output signal in the
−8 steady-state condition, the measurement covariance matrix has been
−12 chosen to be
0 50 100 150 200 250 300 350 400

R  diag0.00242 ; 0.00212 ; 0.00272  rad2 ∕s2 


4
0
−4
and the associated process noise covariance matrix to be
−8
−12 Q  diag10−4 2 ; 10−4 2 ; 10−4 2 ; 3 × 10−3 2 
0 50 100 150 200 250 300 350 400
× rad2 ∕s4 ; rad2 ∕s4 ; rad2 ∕s4 ; m2 ∕s2 
4
0 The estimated value of roffz is shown in Fig. 13. It can be seen that the
−4 estimation of the offset in the zb direction converges to the value of
−3 m. To verify the convergence of the UKF, the
z ≈ 1.5 × 10
roff
−8
−12
0 50 100 150 200 250 300 350 400 square root of the last diagonal element of the error covariance matrix
has been plotted. In fact, if the filter operates correctly, this value in
Downloaded by UNIV OF CALIFORNIA - SANTA CRUZ on June 3, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.60380

Fig. 11 Experimental results: gb components during transversal time should be as close as possible to zero, as shown in Fig. 13.
balancing procedure. The energy conservation method proposed in Sec. IV has been
implemented after the mass balancing procedure, to further verify the
effectiveness of the balancing method. Before the mass balancing
The discrepancies between experimental and simulation results are procedure, an angular velocity has been given to the system and the
particularly evident after t  250 s. In fact, the work of the kinetic rotational energy has been recorded. The procedure is
unmodeled dissipative torque acting on the spacecraft simulator, such repeated after the balancing procedure. Figure 14 shows that the
as aerodynamic drag, causes the system to lose energy and, variation in amplitude of the kinetic energy term is drastically
consequently, to reduce its angular velocity ω. For this reason, as reduced after the mass balancing procedure; thus, it is reasonable to
shown in Fig 12, the experimental value kωg k decreases in time conclude that the proposed mass balancing procedure has produced a
instead of converging to a constant. Another difference between significant reduction of the offset between the center of mass and the
experimental and simulation results appears at the beginning of center of rotation. Indeed, the standard deviation has changed from
the automatic balancing procedure, where the simulated and σ unbalance  2.8223 × 10−4 J to σ balance  8.2474 × 10−5 J, which is
experimental systems, under the same control law, react slightly a reduction of ≈71%. As expected, the reduction in the standard
different. Such difference may be caused by unmodeled dynamics. deviation of the kinetic energy is less than the one obtained in the
For example, the inner-loop control that adjusts the position of simulation. This is because all the measurements from the IMU
the moving masses to the desired location is not modeled in relative to angular velocity and gravity field are affected by noise that
the simulation. Nevertheless, these discrepancies do not affect the has not been modeled in the simulation. Furthermore, other
effectiveness of the balancing procedure because, in both cases, the dissipative torques that are difficult to simulate affect the spacecraft
angular velocity ωp → 0, as deduced analytically in Eq. (22). simulator. The aerodynamic drag is one of the major dissipative
In the second step of the automatic balancing procedure, to torques present during laboratory experimentation. The presence of
compensate the offset in the zb axis, an unscented Kalman filter is different wires and components generates an irregular shape that,
used to estimate the unknown offset. In the experimental validation, exposed to the air flow during the rotational motion, produces a
the UKF sample time has been set equal to onboard computer sample dissipative aerodynamic torque that is difficult to model.
time Ts  0.05 s. The initial error covariance matrix P0 has been Nevertheless, by considering low value of angular velocity, the
chosen as magnitude of the dissipative torque can be considered very small. The
presence of these unmodeled dynamics degrades the estimation
performance. For this reason, the reduction in the standard deviation
P0  diag0.22 ;0.22 ;0.22 ;2 × 10−3 2 rad2 ∕s2 ;rad2 ∕s2 ;rad2 ∕s2 ;m2 
of the kinetic energy during the laboratory experimentation is less
with respect to the simulated one shown in Fig. 8. In particular, by
2.4
1.8 −3
x 10
1.2 2.5
0.6
0 2
−0.6
0 50 100 150 200 250 300 350 400
1.5

2.4
1
1.8
1.2
0.6 0.5
0
−0.6 0
0 50 100 150 200 250 300 350 400

−0.5
2.4
1.8
−1
1.2
0.6
0 −1.5
−0.6
0 50 100 150 200 250 300 350 400
−2
0 5 10 15 20 25 30 35 40
Fig. 12 Experimental and simulation results: Solid lines are
experimental data; dashed lines are simulated data of the magnitude of Fig. 13 Experimental results: roff
z estimation using UKF during vertical
ω, ωp , and ωg during transversal balancing procedure. balancing procedure.
206 CHESI ET AL.

[5] Peck, M. A., and Cavender, A. R., “Airbearing-Based Testbed for


−3
x 10
3
Momentum Control Systems and Spacecraft Line of Sight,” 13th Annual
2.8 AAS/AIAA Spaceflight Mechanics Meeting, American Astronautical
Society Paper 03-127, Feb. 2003.
2.6 [6] Romano, M., and Agrawal, B. N., “Acquisition, Tracking and Pointing
Control of the Bifocal Relay Mirror Spacecraft,” Acta Astronautica,
2.4 Vol. 53, No. 4, 2003, pp. 509–519.
doi:10.1016/S0094-5765(03)80011-5
2.2 [7] Kim, J. J., and Agrawal, B. N., “Automatic Mass Balancing of
Air-Bearing-Based Three-Axis Rotational Spacecraft Simulator,”
2 Journal of Guidance, Control, and Dynamics, Vol. 32, No. 3, 2009,
pp. 1005–1017.
1.8
doi:10.2514/1.34437
1.6
[8] Prado, J., Bisiacchi, G., Reyes, L., Vicenta, E., Contreras, F., Mesinas,
M., and Juarez, A., “Three-Axis Air-Bearing Based Platform for
1.4 Small Satellite Attitude Determination and Control Simulation,”
Journal of Applied Research and Technology, Vol. 3, No. 3, Dec. 2005,
1.2 pp. 222–237.
[9] Prado, J., and Bisiacchi, G., “Dynamic Balancing for a Satellite Attitude
1 Control Simulator. Instrumentation and Development,” Journal of the
Downloaded by UNIV OF CALIFORNIA - SANTA CRUZ on June 3, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.60380

0 10 20 30 40 50
Mexican Society of Instrumentation, Vol. 4, No. 5, 2000, pp. 76–81.
[10] Hatcher, N. M., and Young, R. N., “Automatic Balancing System for
Fig. 14 Experimental results: kinetic energy before Ebr and after Ear the Use on Frictionlessly Supported Attitude-Controlled Test Platforms,”
balancing procedure. NASA TN-D-4426, Langley Research Center, March 1968.
[11] Yang, Y., and Cao, X., “Design and Development of the Small Satellite
Attitude Control System Simulator,” AIAA Modeling and Simulation
observing the energy profile shown in Fig. 14, it is possible to see that Technologies Conference and Exhibit, AIAA Paper 2006-6124,
the energy tends to decrease in time. This decreasing behavior is due Keystone, CO, Aug. 2006.
to the work of the dissipative torques that affect the system. [12] Li, Y., and Gao, Y., “Equations of Motion for the Automatic Balancing
System of 3-DOF Spacecraft Attitude Control Simulator,” Third
International Symposium on Systems and Control in Aeronautics and
VI. Conclusions Astronautics (ISSCAA), IEEE, June 2010, pp. 248–251.
This paper presents an automatic mass balancing system for [13] Jung, D., and Tsiotras, P., “3-DoF Experimental Test-Bed for Integrated
spacecraft simulators, which allows real-time relocation of the center Attitude Dynamics and Control Research,” AIAA Paper 2003-5331,
of mass to the center of rotation using only sliding masses. Compared Aug. 2003.
[14] Schwartz, J. L., and Hall, C. D., “System Identification of a Spherical
with manual balancing, the proposed automatic balancing system can Air-Bearing Spacecraft Simulator,” AAS Paper 2004-122, Feb. 2004.
greatly reduce the labor involved in the balancing process and save [15] Woo, H., Perez, O. R., Chesi, S., and Romano, M., “CubeSat Three
design time. The use of sliding masses as the only actuator makes the Axis Simulator (CubeTAS),” Proceedings of the AIAA Modeling and
implementation relatively easy and the proposed automatic balancing Simulation Technologies Conference, AIAA Paper No. 2011-6271,
system suitable for small-size spacecraft simulators. To overcome the Aug. 2011.
physical limitation of the actuator (i.e., the torque generated by [16] Likins, P. W., Elements of Engineering Mechanics, McGraw–Hill, New
moving masses is constrained in the direction perpendicular to the York, 1973, pp. 438–439.
gravity field), a two-step design is proposed. The performance of the [17] Leine, R. I., and Wouw, N., “Stability and Convergence of Mechanical
proposed automatic balancing method is theoretically justified using Systems with Unilateral Constraints,” Lecture Notes in Applied
and Computational Mechanics, Vol. 36, Springer, New York, 2008,
Lyapunov stability analysis and is demonstrated at simulation level. pp. 145–148.
The experiment conducted on a CubeSat-scale simulator also shows [18] Crassidis, J., and Markley, F., “Unscented Filtering for Spacecraft
the efficacy of the proposed method; however, the presence of Attitude Estimation,” Journal of Guidance, Control, and Dynamics,
unmodeled dynamics (e.g., aerodynamic drag) and different sources Vol. 26, No. 4, 2003, pp. 536–542.
of noise degrades the performance. Further work is needed to doi:10.2514/2.5102
incorporate dissipative torques and to analyze the effect of various [19] Julier, J., and Uhlmann, J. K., “A New Extension of the Kalman Filter to
noises to improve the accuracy. Nonlinear Systems,” Signal Processing, Sensor Fusion, and Target
Recognition VI, Vol. 3068, 1997, pp. 182–193.
[20] Matthew, C. V., Schwartz, J. L., and Christopher, D. H., “Unscented
References Kalman Filtering for Spacecraft Attitude State and Parameter
[1] Kim, B. M., Velenis, E., Kriengsiri, P., and Tsiotras, P., “Designing a Estimation,” Proceedings of the AAS/AIAA Space Flight Mechanics
Low-Cost Spacecraft Simulator,” IEEE Control Systems, Vol. 34, No. 4, Conference, AAS Paper 04-115, 2005.
Aug. 2003, pp. 26–37. [21] Sekhavat, P., Gong, Q., and Ross, I. M., “NPSAT1 Parameter Estimation
[2] Schwartz, J. L., Peck, M. A., and Hall, C. D., “Historical Review of Air Using Unscented Kalman Filtering,” Proceedings of the American
Bearing Spacecraft Simulators,” Journal of Guidance, Control, and Control Conference, IEEE, Piscataway, NJ, July 2007, pp. 4445–4451.
Dynamics, Vol. 26, No. 4, 2003, pp. 513–522. [22] Simon, D., Optimal State Estimation: Kalman, H ∞ , and Nonlinear
doi:10.2514/2.5085 Approaches, Wiley, Hoboken, NJ, 2006, pp. 403–407.
[3] Smith, G. A., “Dynamic Simulators for Test of Space Vehicle Attitude [23] Kang, W., and Barbot, J.-P., “Discussions on Observability and
Control System,” Proceedings of the Conference on the Role of Invertibility,” Seventh IFAC Symposium on Nonlinear Control Systems,
Simulation in Space Technology, Part C, Virginia Polytechnic Inst. and Aug. 2007.
State Univ., Blacksburg, VA, Aug. 1964. [24] Chesi, S., Pellegrini, V., Gong, Q., Cristi, R., and Romano, M.,
[4] Fullmer, R. R., “Dynamic Ground Testing of the Skipper Attitude “Automatic Mass Balancing for Small Three-Axis Spacecraft Simulator
Control System,” 34th AIAA Aerospace Science Meeting and Exhibit, Using Sliding Masses Only,” Advances in the Astronautical Sciences,
AIAA Paper, 1996-0103, Jan. 1996. Vol. 142, No. 3, 2012, pp. 2547–2562.

You might also like