Sie sind auf Seite 1von 22

Rigid-Body

Attitude Control
USING ROTATION MATRICES FOR CONTINUOUS,
SINGULARITY-FREE CONTROL LAWS

30 IEEE CONTROL SYSTEMS MAGAZINE » JUNE 2011 1066-033X/11/$26.00©2011IEEE


R
igid-body attitude control is motivated by aero- Motivated by the desire to represent attitude both glob-
space applications that involve attitude maneu- ally and uniquely in the analysis of rigid-body rotational
vers or attitude stabilization. In addition to space- motion, this article uses orthogonal matrices exclusively to
craft and atmospheric flight vehicles, applications represent attitude and to develop results on rigid-body atti-
of rigid-body attitude control arise for underwater tude control. An advantage of using orthogonal matrices is
vehicles, ground vehicles, and robotic systems. The set of at- that these control results, which include open-loop attitude
titudes of a rigid body is the set of 3 3 3 orthogonal matri- control maneuvers and stabilization using continuous
ces whose determinant is one. This set is the configuration feedback control, do not require reinterpretation on the set
space of rigid-body attitude motion; however, this configu- of attitudes viewed as orthogonal matrices. The main objec-
ration space is not Euclidean. Since the set of attitudes is not tive of this article is to demonstrate how to characterize
a Euclidean space, attitude control is typically studied using properties of attitude control systems for arbitrary attitude
various attitude parameterizations [1]–[8]. These parame- maneuvers without using attitude parameterizations.
terizations can be Euclidean, as in the case of Euler angles,
which lie in R3, or non-Euclidean, as in the case of quaterni- DESCRIPTIONS OF RIGID-BODY
ons, which lie in the non-Euclidean three-sphere S3. ATTITUDE CONFIGURATIONS
Regardless of the choice of parameterization, all param- The attitude of a rigid body can be modeled by a linear
eterizations [1], [2] fail to represent the set of attitudes both transformation between a reference frame and a body-
globally and uniquely. For example, although Euler angles fixed frame that preserves the distance between each pair
can represent every attitude, this representation
ntation of material points in the body and the handedness
is not unique at certain attitudes. In fact, the of coor
coordinate frames [1], [2], [6], [10]–[13].
time derivatives of the Euler angles at thesee Assuming
Assu each frame is defined by three
attitudes cannot represent every possible orthogonal
ort unit vectors ordered accord-
angular velocity. The term gimbal lock is ing
in to the right-hand rule, the attitude of
used to describe this mathematical sin- the
th rigid body, as a linear transforma-
gularity. Along the same lines, quaterni- tion,
ti is represented by a 3 3 3 matrix
ons can represent all possible attitudes that
th transforms a vector resolved in the
and angular velocities, but this repre- body-fixed
b frame into its representation
sentation is not unique; see “Representa- resolved
re in the reference frame. The three
tions of Attitude” for more details. Since column
colu vectors of the matrix represent the
parameterizations such as Euler angles and nd three orthogonal unit basis vectors of the
quaternions are unable to represent the set of at- body-fixed frame resolved in the reference
body-fix
titudes both globally and uniquely, results obtained frame. Likewise, the three rows of the matrix repre-
frame Likew
with these parameterizations have to be reinterpreted on sent the three orthogonal unit basis vectors of the reference
the set of attitudes described by orthogonal matrices. Ne- frame resolved in the body frame. The resulting matrix,
glecting this analysis can result in undesirable behavior referred to as a rotation matrix, represents the physical atti-
such as unwinding [9], which refers to the unstable be- tude of the rigid body. The set of all rotation matrices is the
havior of a closed-loop attitude control system. In par- special orthogonal group of rigid rotations in R3, which is
ticular, for certain initial conditions, the trajectories of denoted by SO 1 3 2 ; see “Attitude Set Notation” for a defini-
the closed-loop system start close to the desired attitude tion of these and additional related sets.
equilibrium in state space and yet travel a large distance In rigid-body pointing applications, the rotation about
in the state space before returning to the desired attitude. the pointing direction is irrelevant since these rotations
This closed-loop trajectory is analogous to the homo- do not change the direction in which the body points.
clinic orbit observed in a simple pendulum that swings Thus, while rotation matrices R [ SO 1 3 2 can be used to
360° when perturbed from its inverted position at rest. Un- model the attitude of a rigid body, not all the information
winding can result in wasted control effort by causing the available in the rotation matrix is required for pointing
rigid body to perform a large-angle slew maneuver when applications. In this case, a reduced representation of the
a small-angle slew maneuver in the opposite rotational di- rotation matrix can be used for pointing applications. An
rection is sufficient to achieve the objective; see “Pitfalls of example of such a reduced representation is the reduced-
Using Quaternion Representations for Attitude Control” attitude vector [14]. To develop the reduced-attitude
for a description and illustration of unwinding in attitude vector representation of a rigid body, suppose b [ S2,
control using quaternions. where S2 is the unit sphere in three-dimensional space, is
a unit vector that denotes a fixed pointing direction speci-
© SHANNON MASH

fied in the reference frame. Then, in the body-fixed frame,


Digital Object Identifier 10.1109/MCS.2011.940459
this direction vector can be expressed as G ! R^b [ S2,
Date of publication: 16 May 2011 where R^ denotes the transpose of the rotation matrix R

JUNE 2011 « IEEE CONTROL SYSTEMS MAGAZINE 31


Representations of Attitude
igid body attitude is often represented using three or four
R parameters. Unit quaternions and the axis-angle represen-
tation use four parameters to represent the attitude. Three-pa-
Table S1 Properties of attitude representations.

rameter representations of attitude include the Euler angles as Attitude Representation Global? Unique?
Euler angles No No
well as parameters derived from the unit quaternions as in the
Rodrigues parameters No No
case of the Rodrigues parameters and the modified Rodrigues Modified Rodrigues parameters No No
parameters. These three-parameter sets can be viewed as Quaternions Yes No
embedded subsets of R3, thus allowing methods of analysis Axis-angle Yes No
that are suited to the Euclidean space R3. Euler angles are Rotation matrix Yes Yes
kinematically singular since the transformation from their time
rates of change to the angular velocity vector is not globally
defined. The Rodrigues parameters and modified Rodrigues As shown in Table S1, since the unit quaternions are global
parameters are geometrically singular since they are not de- in their representation of attitude, they are often used in practi-
fined for 180° of rotation. Therefore, continuous control laws cal applications to represent rigid-body attitude. However, note
using these three-parameter representations cannot be glob- that the map from the space S3 of unit quaternions to the space
ally defined, and, as such, these representations are limited to SO 1 3 2 of rotations is not unique. In particular, this map is a two-
local attitude maneuvers. fold covering map, where each physical attitude R [ SO 1 3 2 is
Table S1 presents various attitude representations and represented by a pair of antipodal unit quaternions 6q [ S3. If
their key properties. While the quaternions, axis-angle vari- not carefully designed, quaternion-based controllers may yield
ables, and rotation matrices can represent all attitudes of a undesirable phenomena such as unwinding [9]; see “Pitfalls
rigid body, only rotation matrices can represent all attitudes of Using Quaternion Representations for Attitude Control” for
uniquely. more details.

representing the attitude of the rigid body. We refer to the lems using continuous feedback control. In particular, we
vector G as the reduced-attitude vector of the rigid body. seek continuous attitude control laws that modify the
Thus, a reduced-attitude vector describes a direction b dynamics of the spacecraft such that a specified attitude or
fixed in the reference frame by a unit vector resolved in reduced attitude is rendered asymptotically stable. For full-
the body frame. In pointing applications, a pointing direc- attitude control, this continuous feedback function is
tion b [ S2 in the reference frame can be expressed by the defined in terms of the full rigid-body attitude and angular
reduced attitude G corresponding to b, and, hence, the velocity, while, for reduced-attitude control, this feedback
reduced attitude can be used to study pointing applica- function is defined in terms of the reduced attitude and
tions for a rigid body. angular velocity.

FULL- AND REDUCED-ATTITUDE CONTROL LOCAL AND GLOBAL


Rigid-body attitude control problems can be formulated ATTITUDE CONTROL ISSUES
in terms of attitude configurations defined in SO 1 3 2 or in Since attitude control problems are nonlinear control prob-
terms of reduced-attitude configurations defined in S2. lems, these problems can be categorized into local attitude
The former case involves control objectives stated directly control issues and global attitude control issues. In brief,
in terms of the full attitude of the rigid body, while the local control issues address changes in the rigid-body atti-
latter case involves control objectives stated in terms of tude and angular velocity that lie in an open neighborhood
pointing the rigid body. In particular, the latter case is of the desired attitude and angular velocity. Local attitude
used when the control objectives can be formulated in control problems typically arise when the rigid body has a
terms of pointing a body-fixed imager, antenna, or other limited range of rotational motion. In this case, linear con-
instrument in a specified direction in the reference frame, trol methods or local nonlinear control methods based on a
where rotations about that specified body-fixed direction convenient attitude representation may suffice, so long as
are irrelevant. no singularities occur within the local range of attitude
We subsequently study several full-attitude and motion of interest.
reduced-attitude control problems. In each case, we begin In contrast to local attitude control, global attitude
by treating open-loop attitude control problems defined by control issues arise when arbitrary changes in the
specifying initial conditions, terminal conditions, and a rigid-body attitude and angular velocity are allowed.
maneuver time. Then we treat attitude-stabilization prob- In this case, no a priori restrictions are placed on the

32 IEEE CONTROL SYSTEMS MAGAZINE » JUNE 2011


possible rotational motion, and nonlinear methods are FRAMEWORK FOR
required to address these rotational motions. Global DESCRIBING ATTITUDE MOTION
attitude control issues are motivated by applications The space of rigid-body rotations SO 1 3 2 is a Lie group [34],
to highly maneuverable rigid-body systems that may [35]; it has an algebraic group structure based on matrix
undergo extreme attitude changes such as rigid-body multiplication as the group operation and a differential
tumbling. geometric structure as a compact smooth manifold that
enables the use of calculus. Alternatively, the Lie group
COMMENTS ON THE SO 1 3 2 of rigid-body rotations can be viewed as the set of
ATTITUDE CONTROL LITERATURE isometries, or length-preserving transformations on R3
The literature on attitude control can be divided into two that leave the origin fixed and preserve the orientation, that
categories. In the first category, attitude control is studied is, handedness, of frames. Unlike SO 1 3 2 , the space of rigid-
within a geometric framework, as presented in this article, body reduced-attitudes S2 is not a Lie group. However,
by viewing rigid-body attitude as an element of SO 1 3 2 or S2 each element in SO 1 3 2 , as an orthogonal matrix, is a bijec-
[14]–[27]. In the second category [3]–[8], [28], [29], attitude tive mapping from S2 to S2.
control is studied using representations of attitude in R3 or
S3 ( R4 [1], [2]. Attitude Description Using Rotation Matrices
Almost-global attitude stabilization of a rigid body, We now describe the properties of SO 1 3 2 . Every rotation
in contrast to global attitude stabilization, is motivated matrix R [ SO 1 3 2 satisfies
by the fact that there exists no continuous time-invari-
R^R 5 I 5 RR^, det 1 R 2 5 1, (1)
ant feedback that globally asymptotically stabilizes a
single equilibrium attitude [16]–[18]. This property is where I denotes the 3 3 3 identity matrix, which is also
related to the fact that every continuous time-invariant the identity element of the group SO 1 3 2 . This representa-
closed-loop vector field on SO 1 3 2 and S2 has more than tion of the attitude is global and unique in the sense that
one closed-loop equilibrium, and hence, the desired each physical rigid-body attitude corresponds to exactly
equilibrium cannot be globally attractive [9]; see “The one rotation matrix. If Rd [ SO 1 3 2 represents the desired
Impossibility of Global Attitude Stabilization Using attitude of a rigid body, then attitude control objectives
Continuous Time-Invariant Feedback” for more details. can be described as rotating the rigid body so that its atti-
While [24] treats local attitude stabilization and H` atti- tude R is the same as the desired attitude, that is, R 5 Rd.
tude performance issues, [19]–[23] present results for The tangent space at an element in the Lie group
various almost-global stabilization problems for the 3D SO 1 3 2 consists of vectors tangent to all differentiable
pendulum; the 3D pendulum has the same rigid-body curves t S R 1 t 2 [ SO 1 3 2 passing through that element.
attitude dynamics as in this article, but the moment due Thus, the tangent space at an element in SO 1 3 2 is the
to uniform gravity is also included [30], [31]. This same vector space of all possible angular velocities that a rigid
perspective is used in [25] and [26], where results on body can attain at that attitude. This vector space is iso-
almost-global tracking of desired attitude trajectories morphic to the tangent space at the identity element
are presented. Rigid-body reduced-attitude control is R 5 I, which is defined as the Lie algebra so 1 3 2 of the Lie
studied in [14], [15], [22], [23], [27], and [32]. group SO 1 3 2 . The algebraic structure of so 1 3 2 is obtained
An alternative approach to attitude control based on from the defining equation R 1 t 2 ^R 1 t 2 5 I 5 R 1 t 2 R 1 t 2 ^. Let
#
the geometry of SO 1 3 2 or S2 is to use representations of R 1 t 2 [ SO 1 3 2 be a curve such that R 1 0 2 5 I and R 1 0 2 5 S,
attitude in R3 or S3 ( R4 [1], [2], such as Euler angles, Euler where S [ so 1 3 2 . Differentiating R 1 t 2 ^R 1 t 2 5 I and
parameters, or quaternions, Rodrigues parameters, and evaluating at t 5 0, we see that S^1 S 5 0, that is, S is
modified Rodrigues parameters [3]–[8], [15], [28], [29], skew symmetric. Thus, so 1 3 2 is the matrix Lie algebra of
[33]. While these representations can be used for local atti- 3 3 3 skew-symmetric matrices.
tude control, global results obtained by using this approach Each element of SO 1 3 2 can be represented by the matrix
must be interpreted on the configuration space SO 1 3 2 of exponential [34]
the physical rigid body. Indeed, the global dynamics of the
1 2 1 3 c
parameterized closed-loop attitude motion on R3 or S3, exp 1 S 2 5 I 1 S 1 S 1 S 1 , (2)
2! 3!
such as global or almost global asymptotic stability, do not
necessarily imply that these closed-loop properties also where S is in the Lie algebra so 1 3 2 and exp 1 S 2 is in the Lie
hold for the dynamics of the physical rigid body evolving group. The Lie algebra so 1 3 2 is isomorphic to R3, and this
on the configuration space SO 1 3 2 . Issues related to contin- vector space isomorphism is given by
uous attitude control using these alternative representa-
tions of attitude are described in “Representations of 0 2 s3 s2
Attitude” and “Pitfalls of Using Quaternion Representa- S 5 £ s3 0 2 s1 § [ so 1 3 2 , (3)
tions for Attitude Control.” 2 s2 s1 0

JUNE 2011 « IEEE CONTROL SYSTEMS MAGAZINE 33


Pitfalls of Using Quaternion Representations for Attitude Control

Q
uaternion representations of attitude are often used in at- In the case of quaternions, continuous, time-invariant state-
titude control and stabilization to represent rigid-body at- feedback control using these representations can give rise to
titude. Quaternions do not give rise to singularities, but they regions of attraction and repulsion in the neighborhood of
double cover the set of attitudes SO 1 3 2 in the sense that each q [ S3 and 2q [ S3, respectively, where q and 2q represent
attitude corresponds to two different quaternion vectors. Spe- the same desired attitude in SO 1 3 2 . Starting from q, we can
cifically, a physical attitude R [ SO 1 3 2 is represented by a pair choose an initial angular velocity such that orbits connecting
of antipodal quaternions 6q [ S3. these regions of repulsion and attraction arise. These orbits
An implication of this nonunique representation is that run from a neighborhood of q to a neighborhood of 2q in S3,
closed-loop properties derived using quaternions might with a terminal velocity that is close to zero. The correspond-
not hold for the dynamics of the physical rigid body on ing orbit in SO 1 3 2 obtained by projection as shown in Figure
SO 1 3 2 3 R3. Thus, almost global asymptotic stability of an S1 starts and ends close to the desired attitude. However, in
equilibrium in the quaternion space S3 3 R3 does not guar- the quaternion space, since 2q has a region of repulsion, the
antee almost global-asymptotic stability of the correspond- trajectory is repelled from 2q, eventually converging to q. In
ing equilibrium in SO 1 3 2 3 R3. Indeed, a property such as SO 1 3 2 , the corresponding trajectory then follows a homoclin-
local asymptotic stability of the closed-loop vector field on ic-like orbit starting close to the desired attitude equilibrium
S3 3 R3 holds on SO 1 3 2 3 R3, only if the property holds for with almost zero angular velocity, gaining velocity as the rigid
the projection of the vector field from S3 3 R3 to the space body moves along the orbit, and eventually converging to the
SO 1 3 2 3 R3; see Figure S1. desired equilibrium. Thus, the rigid body unwinds needlessly
If the projection of the closed-loop vector field from before converging to the desired attitude. The presence of
S3 3 R3 onto SO 1 3 2 3 R3 is neglected during control design, these homoclinic-like orbits in SO 1 3 2 indicates that, although
then undesirable phenomena such as unwinding can occur the desired closed-loop attitude equilibrium is attractive, it is
[9]. In unwinding, a closed-loop trajectory in response to cer- not Lyapunov stable.
tain initial conditions can undergo a homoclinic-like orbit that To illustrate unwinding, suppose the quaternion q [ S3 rep-
starts close to the desired attitude equilibrium. Unwinding resents the attitude of a rigid body, and let v [ R3 represent its
can occur in all continuous closed-loop attitude control sys- angular velocity vector resolved in the body-fixed frame. We ex-
tems that are designed using global parameterizations that press q 5 1 q0, qv 2 , where 1 q0, qv 2 [ R 3 R3. Then, the attitude
are nonunique in their representation of SO 1 3 2 ; see “Repre- dynamics of the rigid body in the quaternion space are given by
sentations of Attitude” for examples of global parameteriza-
#
tions that are nonunique in its representation of SO 1 3 2 . Jv 5 J v 3 v 1 u,
# 1
q 5 2 qTv v,
2
# 1
qv 5 1 qI 1 qv3 2 v,
Tangent A 3
2
(Quaternion Space)
Plane
where J is the moment of inertia matrix, I is the 3 3 3 iden-
Projection from
Quaternion to Physical tity matrix, and u [ R3 is the applied control input. Next,
B Rotation Space consider the PD-type control law u 1 q, v 2 ! 2kpqv 2 kd v,
SO(3) where kp, kd . 0. It can be shown that qe 5 1 1, 0, 0, 0 2 and
A′
B′ (Rotation Space) v e 5 1 0, 0, 0 2 is an equilibrium of the resulting closed-loop
vector field. In the physical rotation space, this equilibrium cor-
responds to the attitude Re 5 I and v e 5 0. Furthermore, the
closed-loop vector field yields almost global asymptotic sta-
Tangent Plane bility for the equilibrium 1 qe, v e 2 [ S3 3 R3. However, almost
global asymptotic stability of 1 qe, v e 2 [ S3 3 R3 does not im-
FIGURE S1 Projection of the vector field from S3 3 R3 to ply almost global asymptotic stability for the corresponding
SO 1 3 2 3 R3. Since orthogonal matrices represent the set of equilibrium 1 Re, v e 2 [ SO 1 3 2 3 R3.
attitudes globally and uniquely, results obtained using alterna-
To illustrate this failure, we choose J 5 diag 1 1, 1, 1 2 kg-m2
tive parameterizations of SO 1 3 2 have to be reinterpreted on the
set of attitudes described by orthogonal matrices. Neglecting and 1 kp, kd 2 5 1 0.1, 0.237 2 . Furthermore, choose the initial atti-
this analysis can result in undesirable behavior such as unwind- tude q 1 0 2 5 qe and initial angular velocity v 1 0 2 5 3 360/p 0 0 4 °/s.
ing [9]. To interpret the results on SO 1 3 2 , the closed-loop trajec- The quaternion element q0 1 t 2 is shown in Figure S2(a), and
tory is projected from the parametric space onto SO 1 3 2 , where
the angular velocity is shown in Figure S2(b). Furthermore,
the properties of the resulting closed-loop on SO 1 3 2 are ana-
lyzed. This figure illustrates the projection of a trajectory from Figure S2(c) shows the error in the attitude in SO 1 3 2 using the
the space of quaternions S3 onto SO 1 3 2 . eigenangle

34 IEEE CONTROL SYSTEMS MAGAZINE » JUNE 2011


1
U 1 t 2 5 cos 21 1 trace 1 R 1 t 2 2 2 1 2 [ 3 0, 180 4 ,
2
1

which is a scalar measure of the error between R and I. Note


that since Re 5 I, it follows that U 5 0° for R 5 Re. For the cho- 0.5
sen initial conditions, the eigenangle satisfies U 1 0 2 5 0°.
As shown in Figure S2(a) and (b), the trajectory in S3

q0
0
approaches 2qe at t 5 17 s with a small angular velocity
of magnitude 0.03 °/s and is then repelled away, eventu- q0 ≈ −0.999
ally converging to the desired attitude qe. This closed-loop −0.5
behavior is similar to the response of a simple pendulum
with angular velocity damping that is swung up to the un- −1
stable inverted equilibrium starting from the stable hanging 0 20 40 60 80 100
Time (s)
equilibrium, where the pendulum eventually converges to
(a)
the stable hanging equilibrium. Since 1 qe, 0 2 [ S3 3 R3 is
120
a stable equilibrium and 1 2qe, 0 2 [ S3 3 R3 is an unstable
equilibrium of the closed-loop dynamics, the closed-loop 100
trajectory in S3 3 R3 shows a closed-loop response similar
to the simple pendulum swinging up with angular velocity 80
damping. ||ω|| (°/s)
60
However, when the closed-loop trajectory in S3 3 R3 is pro- ||ω|| ≈ 0.03 °/s
jected onto SO 1 3 2 3 R3, parts (b) and (c) of Figure S2 show 40
that the corresponding trajectory in SO 1 3 2 3 R3 unwinds. Thus,
the closed-loop trajectory approaches the desired attitude Re 20
1 U 1 t 2 < 2.75°) with a small angular velocity 1 7 v 1 t 2 7 < 0.03 °/s)
0
at t 5 17 s, and yet instead of converging to the equilibrium 0 20 40 60 80 100
attitude Re, the trajectory is repelled away from 1 Re, 0 2 with Time (s)
an increasing angular velocity. Indeed, U 1 t 2 in Figure S2(c) (b)
increases to 180° at t < 50 s, eventually converging to zero. 200
Thus, starting at t 5 17 s, we observe a homoclinic-like orbit
in SO 1 3 2 3 R3 at 1 Re, 0 2 , which violates the stability of the 150
equilibrium. The presence of this homoclinic-like trajectory indi-
cates that, while the desired equilibrium is globally attractive in
Θ (°)

100
SO 1 3 2 3 R3, it is not Lyapunov stable and, hence, not asymp-
totically stable in SO 1 3 2 3 R3.
In the above example, the controller u 1 q, v 2 ! 2kp qv 2 kd v 50 Θ ≈ 2.75°
fails to achieve asymptotic stability on the physical space
SO 1 3 2 3 R3 since it does not satisfy u 1 q, v 2 5 u 1 2q, v 2 . Thus,
0
while q and 2q in the quaternion space represent the same 0 20 40 60 80 100
physical attitude in SO(3) the control input u 1 q, v 2 2 u 1 2q, v 2 . Time (s)
Therefore, for each 1 R, v 2 [ SO 1 3 2 3 R3, where R 2 Re, the (c)
controller yields two distinct control inputs. Hence, the projec-
tion of the closed-loop vector field from S3 3 R3 to SO 1 3 2 3 R3 FIGURE S2 Closed-loop response of the rigid body as observed
yields a multivalued map. Thus, while the quaternion-based in S3 and SO 1 3 2 . As shown in (b) and (c), the rigid body
controller results in a closed-loop system that is a continuous approaches the desired attitude 1 U 5 2.75°) with an angular
velocity whose magnitude is approximately 0.03 °/s. Thus,
vector field on S3 3 R3, the projected closed-loop system on
from asymptotic stability, we expect the trajectory to converge
the physical space SO 1 3 2 3 R3 is not a vector field. Since the to the desired equilibrium 1 Re, 0 2 . On the contrary, the trajec-
projected closed loop on SO 1 3 2 3 R3 is not a vector field, we tory is repelled from 1 Re, 0 2 , with an increasing angular veloc-
cannot conclude that closed-loop properties on S3 3 R3, such ity. This homoclinic-like orbit occurs since q in (a) approaches
as asymptotic stability and almost global-asymptotic stability, the unstable attitude 2qe and is subsequently repelled away,
converging eventually to qe. However, in the configuration
imply the same on SO 1 3 2 3 R3.
space SO 1 3 2 , both qe and 2qe correspond to the same attitude
In practical applications, unwinding is often resolved by and hence the homoclinic-like orbit. This phenomenon is
changing the sign of the quaternion vector as the closed-loop termed unwinding.

JUNE 2011 « IEEE CONTROL SYSTEMS MAGAZINE 35


trajectory passes from one hemisphere to the other. Indeed, the the quaternion-based controller in S3 3 R3. Stability analysis
only way to define a closed-loop vector field on SO 1 3 2 3 R3 for discontinuous closed-loop systems requires set-valued dif-
for a continuous time-invariant, state-feedback control- ferential equations [41].
ler using quaternions is to choose a controller that satisfies In contrast to analysis using quaternions, none of the above
u 1 q, v 2 5 u 1 2q, v 2 . However, changing the sign of the qua- issues occur for closed-loop vector fields that are defined di-
ternion vector as the closed-loop trajectory passes from one rectly on SO 1 3 2 3 R3 by means of rotation matrices. This prop-
hemisphere to the other results in a discontinuous closed-loop erty is due to the fact that, unlike quaternions, rotation matri-
vector field on S3 3 R3. Note that almost global asymptotic ces represent rigid-body attitudes both uniquely and globally.
stability of the desired attitude for the continuous controller in This uniqueness property is one of the primary motivations for
S3 3 R3 does not imply almost global asymptotic stability of studying attitude-control issues on SO 1 3 2 3 R3 in terms of ro-
the desired attitude for the discontinuous implementation of tation matrices.

where s 5 3 s1 s2 s3 4 ^ [ R3. This isomorphism is expressed For the fixed reference frame direction b [ S2, dif-
as S 5 s 3 . The reason for the superscript is related to the ferentiating the reduced attitude G 5 R^b with respect
fact for every a, b [ R3, 3 a 3 4 b 5 a 3 b. If we identify to time and substituting the kinematics equation (5) for
S [ so 1 3 2 as s 5 aa, where a [ S2 and a [ R, then this iso- the full attitude R [ SO 1 3 2 yields the kinematics for the
morphism gives Euler’s theorem, where a [ S2 is the axis reduced attitude. Since v 3 is skew symmetric, it fol-
of rotation and a [ R is the angle, corresponding to the lows that
rotation matrix exp 1 S 2 [ SO 1 3 2 . Using this isomorphism,
# #
we can express the matrix exponential of S [ so 1 3 2 by G 5 R^b 5 1 v 3 2 ^R^b 5 1 v 3 2 ^G 5 2 v 3 G 5 G 3 v. (6)
Rodrigues’s formula
This kinematics equation is consistent with the fact that the
exp 1 S 2 5 I 1 a 3 sin 1 a 2 1 1 a 3 2 2 1 1 2 cos 1 a 22 . (4) reduced-attitude vector G evolves on S2. Indeed, differenti-
ating 7G 7 2 ! G^G with respect to time and substituting (6)
Reduced-Attitude Description yields
Using Unit Vectors in R3
Let G ! R^b denote the unit vector b that is fixed in the d ^ #
1 G G 2 5 2G^ G 5 2G^ 1 G 3 v 2 5 2v ^ 1 G 3 G 2 5 0.
reference frame, resolved in the body-fixed frame. Then dt
G [ S2 describes the reduced attitude of a rigid body. Let
Gd 5 R ^
d b [ S represent the desired pointing direction of
2
Therefore, since G 1 0 2 [ S2, hence G 1 t 2 [ S2 for every t $ 0.
an imager, antenna, or other instrument. Pointing in Note that the component of the angular velocity vector v
terms of reduced-attitude control objectives can then be along the vector G does not affect the kinematics (6).
described as rotating the rigid body so that its reduced-
attitude vector G is aligned with the desired reduced-atti- Attitude Dynamics
tude vector Gd, that is G 5 Gd, where both reduced-attitude We now define the attitude dynamics of a rigid body
vectors are resolved in the body-fixed frame. Rotation with respect to an inertial reference frame. The rate of
about the unit vector b does not change the reduced-atti- change of the rotation matrix R [ SO 1 3 2 depends on the
tude defined by the direction of b. Therefore, all full-body angular velocity through the kinematics given by (5).
attitudes that are related by such a rotation result in the The rate of change of the angular velocity expressed in
same reduced attitude. the body frame is Euler’s equation for a rigid body, which
is given by
Attitude Kinematics #
Jv 5 Jv 3 v 1 u, (7)
We denote the attitude of a rigid body by R [ SO 1 3 2 rela-
tive to a reference frame. Let v [ R3 be the angular veloc- where u is the sum of external moments applied to the body
ity of the body relative to the reference frame, resolved in resolved in the body frame. We assume that the rigid body
the body-fixed frame. Then the full-attitude kinematics is fully actuated, that is, the set of all control inputs that
equation that gives the time rate of change of the rigid- can be applied to the rigid body is a three-dimensional
body attitude is Euclidean space.

#
R 5 Rv 3 , (5) GLOBAL ATTITUDE CONTROL MANEUVERS
We now consider attitude control maneuvers that rotate
where v 3 is a skew-symmetric matrix as given in (3). the rigid body from a specified initial attitude and initial

36 IEEE CONTROL SYSTEMS MAGAZINE » JUNE 2011


angular velocity to a specified terminal attitude and termi-
nal angular velocity, in a specified time period. The atti- Attitude Set Notation
tude control maneuvers are defined by describing the
he following set-theoretic notation is used throughout
control moment input, the resulting attitude, and the
angular velocity as functions of time over the specified
T this article.
• 7 x 7 is the Euclidean norm of x.
time period. We consider fully actuated open-loop control
• S1 5 5 1 x1, x2 2 [ R2: i 1 x1, x2 2 i 5 1 6 is the unit circle.
problems with a control function u: 3 0, T 4 S R3. If the ini-
• S2 5 5 1 x1, x2, x3 2 [ R3: i 1 x1, x2, x3 2 i 5 1 6 is the two-
tial angular velocity is zero and the terminal angular
sphere.
velocity is zero, then the attitude control maneuver is a
• S3 5 5 1 x1, x2, x3, x4 2 [ R4: i 1 x1, x2, x3, x4 2 i 5 1 6 is the
rest-to-rest attitude maneuver.
three-sphere.
• SO 1 3 2 5 5 R [ R333: RRT 5 RTR 5 I, det 1 R 2 5 1 6 is the
Full-Attitude Maneuvers
special orthogonal group.
The attitude maneuver considered here is to transfer the
• so 1 3 2 5 5 U [ R333: UT 5 2U 6 is the space of 3 3 3
initial attitude R0 [ SO 1 3 2 and zero initial angular velocity
skew-symmetric matrices, also the Lie algebra of
to the terminal attitude Rf [ SO 1 3 2 and zero terminal
SO 1 3 2 .
angular velocity in the fixed maneuver time T . 0.
According to Euler’s theorem [1], there exists an axis
a [ S2 and an angle a [ 3 0, 2p 2 that satisfy
3 tude maneuver is required, whereas if they differ by a
eaa 5 Rf R021. (8)
sign, then a is chosen to be an arbitrary direction perpen-
A rest-to-rest attitude maneuver that meets the specified dicular to G0 with a 5 p. A rest-to-rest attitude maneuver
boundary conditions is given by that meets the specified boundary conditions is

R 1 t 2 5 e u 1 t 2 a R0 , G 1 t 2 5 e a u 1 t 2 G0
3 3

5 3 I 1 a 3 sinu 1 t 2 1 1 a 3 2 2 1 1 2 cosu 1 t 2 2 4 R0, (9) 5 3 I 1 a 3 sinu 1 t 2 1 1 a 3 2 2 1 1 2 cosu 1 t 2 2 4 G0, (15)


# #
v 1 t 2 5 u 1 t 2 a, (10) v 1 t 2 5 u 1 t 2 a, (16)
$ # $ #
u 1 t 2 5 u 1 t 2 Ja 1 u 1 t 2 2 1 a 3 Ja 2 , (11) u 1 t 2 5 u 1 t 2 Ja 1 u 1 t 2 2 1 a 3 Ja 2 , (17)

where the rotation angle u: 3 0, T 4 S S1 satisfies where the rotation angle u: 3 0, T 4 S S1 satisfies (12) and
(13).
# To achieve a full-attitude maneuver or a reduced-attitude
u 1 0 2 5 0, u 1 0 2 5 0, (12)
maneuver that is not rest to rest, one possible approach is to
a, 0 # a , p, # select a sequence of three successive rotations. The first phase
u 1T2 5 e u 1 T 2 5 0. (13)
2 2p 1 a, p # a , 2p, is based on selecting a fixed-axis rotation (with the rotation
axis fixed in both the inertial and body frames) that brings the
The rotation angle u 1 t 2 can be chosen as a linear combina- initial angular velocity of the rigid body to rest. The third
tion of sinusoids or polynomials with coefficients selected phase is based on selecting a similar fixed-axis rotation that,
to satisfy the initial and terminal boundary conditions (12) in reverse time, brings the terminal angular velocity of the
and (13). rigid body to rest. The second phase uses the framework pre-
sented in (8)–(13) and (14)–(17) to construct the rest-to-rest
Reduced-Attitude Maneuvers maneuver that transfers the attitude at the end of phase one to
We now consider a reduced-attitude maneuver that trans- the required attitude at the beginning of phase three.
fers an initial reduced attitude G0 [ S2 and zero initial The framework presented in (8)–(13) and (14)–(17) can
angular velocity to a terminal reduced attitude Gf [ S2 and form the basis for planning attitude maneuvers. For example,
zero terminal angular velocity in the fixed maneuver time u 1 t 2 in (9)–(13) and (15)–(17) can be viewed as depending on
T . 0. We follow the strategy of a rotation about an iner- parameter values that can be adjusted to achieve some atti-
tially fixed axis. This axis, normalized to lie in S2, is given tude maneuver objective such as minimum time or mini-
by a 5 G0 3 Gf / 7G0 3 Gf 7 [ S2 assuming G0 and Gf are not mum integrated norm of the control input using optimization
colinear. The angle a [ 3 0, 2p 2 satisfies algorithms. The key is the parameterization of attitude
maneuvers in a form that satisfies the boundary conditions.
cos 1 a 2 5 G^
0 Gf. (14)
Observations on Controllability
If G0 and Gf are colinear, then these vectors are either The constructions given above demonstrate that the fully
equal or differ by a sign; if they are equal, then no atti- actuated rigid body is controllable even over arbitrarily

JUNE 2011 « IEEE CONTROL SYSTEMS MAGAZINE 37


The Impossibility of Global Attitude Stabilization Using Continuous Time-Invariant Feedback
he impossibility of global attitude stabilization using con-
T tinuous time-invariant feedback is illustrated for planar
rotations. The insight obtained for the case of attitude con-
C

figurations on the unit circle S1 is helpful for understanding


the analogous results when the attitude configurations are in
SO 1 3 2 or S2. Consider the attitude stabilization of the arm of
the clock needle shown in Figure S3. We desire to stabilize A B
the needle to the configuration in S1 indicated by point A by
applying a force at the tip of the needle, which is tangent to
the circle. Typical initial configurations are indicated by points
B, C, or D. Each configuration can be moved to configuration D
A by either a clockwise force or a counterclockwise force, with
the tip of the needle defining trajectories in S1 as shown in FIGURE S3 Illustration of the impossibility of global stabili-
Figure S3. This explicit construction of forces indicates con- zation for planar rotations. In this case, the configuration
trollability. However, if we attempt to construct a continuous manifold is the circle S1. To stabilize configuration A, a con-
tinuous force vector field that rotates the needle to configu-
force vector field that rotates the needle to configuration A from
ration A from each configuration or point on the circle is
each configuration or point on the circle, then an additional constructed as shown by the vectors tangential to the
equilibrium (B) is born. This equilibrium (B) is created because circle. However, since this vector field points in the oppo-
on the upper half and lower half of the circle, the force vector site direction in the upper half and lower half of the circle,
fields are nonvanishing and point in opposing directions (either the force vector field must vanish at some point B. The van-
ishing of the vector field at B creates a second equilibrium
clockwise or counterclockwise) toward A. Since the force vec-
in addition to A. In a similar fashion, continuous time-invari-
tor field is continuous, it must vanish at some point B, thereby ant closed-loop vector fields yield multiple closed-loop
creating a second equilibrium in addition to A. equilibria on SO 1 3 2 and S2.

short maneuver times. This observation is true for both tion. The objective in full-attitude stabilization is to select a
full-attitude maneuvers and reduced-attitude maneuvers, feedback control function
thereby providing explicit constructions that demonstrate
controllability. u : SO 1 3 2 3 R3 S R3 (18)
We now show that the constructed maneuvers (8)–
(13) and (14)–(17) are not continuous functions of the that asymptotically stabilizes a desired rigid body attitude
boundary conditions. For example, consider a reduced- equilibrium, while the objective in reduced-attitude stabili-
attitude maneuver where G0 and Gf lie along the north zation is to select a feedback control function
and south pole of S2, respectively. In this case, it follows
u : S2 3 R3 S R3 (19)
from (12) and (13) that a slight perturbation in the ini-
tial condition can result in a control maneuver where that asymptotically stabilizes a desired rigid body reduced-
the reduced attitude evolves through either the west or attitude equilibrium. The desired attitude equilibrium or
the east hemisphere of S2 depending on the perturba- the reduced-attitude equilibrium is assumed to be speci-
tion. An analogous phenomenon occurs for attitude fied, and the desired angular velocity vector is zero.
maneuvers on SO 1 3 2 . Thus, minor changes in the values Since we assume full actuation, linear control theory
of the initial attitude, initial angular velocity, terminal can be applied to the linearization of the rigid-body equa-
attitude, and terminal angular velocity may result in tions of motion around a desired equilibrium to achieve
major changes in the attitude maneuver. This disconti- local asymptotic stabilization. However, no continuous
nuity is not due to the particular approach used for time-invariant feedback controller can globally asymp-
designing open-loop controllers. Rather, this disconti- totically stabilize an equilibrium attitude of a rigid body
nuity is a consequence of the non-Euclidean geometry [9], [16]; see “The Impossibility of Global Attitude Stabili-
of the special orthogonal group SO 1 3 2 and the two- zation Using Continuous Time-Invariant Feedback.” In
sphere S2. fact, if we construct a continuous time-invariant feedback
controller that locally asymptotically stabilizes a desired
GLOBAL ATTITUDE STABILIZATION attitude equilibrium on SO 1 3 2 3 R3 or S2 3 R3, then the
We now study fully actuated attitude stabilization and resulting closed-loop vector field must have more than
present results for both full- and reduced-attitude stabiliza- one equilibrium.

38 IEEE CONTROL SYSTEMS MAGAZINE » JUNE 2011


Since no continuous time-invariant feedback controller Closed-Loop Equilibria
can globally asymptotically stabilize a desired attitude Let 1 Re, v e 2 be an equilibrium of the closed-loop system
equilibrium, we are faced with the following questions: (24)–(25). Equating the right-hand side of (25) to zero yields
1) Since global asymptotic attitude stabilization using a Rev e3 5 0. Since Re is invertible, we obtain v e 5 0. Substitut-
continuous time-invariant feedback controller is ing v e 5 0 in (24) and equating it to zero yields V a 1 Re 2 5 0.
impossible, can we obtain closed-loop stabilization It can be shown [23] that there are four closed-loop equilib-
properties that are nearly global? ria that satisfy V a 1 Re 2 5 0 and v e 5 0.
2) What are the global properties of the resulting closed-
loop system for a continuous time-invariant feedback Proposition 1
controller? 1 Re, 0 2 is an equilibrium of the closed-loop attitude equa-
3) Are there closed-loop performance implications for a tions of motion (24)–(25) if and only if
continuous time-invariant feedback controller?
d Re [ S ! 5 3 e1 e2 e3 4 , 3e1 2 e2 2 e3 4 , 32e1 2 e2 e3 4 ,
R^
FULL-ATTITUDE STABILIZATION 32e1 e2 2 e3 46 . (26)
We now present a proportional and derivative feedback
control structure and characterize the resulting closed-loop That is, 1 Re, 0 2 [ ~, where
vector field. We use this example to answer the above ques-
tions. We consider rigid-body attitude stabilization using
~ ! 51 Rd, 0 2 , 1 Ra, 0 2 , 1 Rb, 0 2 , 1 Rg, 0 26 , (27)
the fully actuated rigid-body equations of motion

#
Jv 5 Jv 3 v 1 u, (20) where
#
R 5 Rv 3 , (21)
Ra ! Rd 3 e 1 2e2 2e3 4 , (28)
where R [ SO 1 3 2 and v [ R3.
Rb ! Rd 3 2e1 2e2 e3 4 , (29)
The full-attitude stabilization objective is to select a con-
tinuous feedback controller u : SO 1 3 2 3 R3 S R3 that Rg ! Rd 3 2e1 e2 2e3 4 . (30)
asymptotically stabilizes the desired attitude equilibrium
given by the desired attitude Rd [ SO 1 3 2 and the angular Proof
velocity v 5 0. The feedback controller is given by Equating V a 1 Re 2 5 0 yields

u 5 2Kvv 2 KpV a 1 R 2 , (22)


a 1 e 1 3 R^ ^ ^
d Ree1 1 a2e2 3 Rd Ree2 1 a3e3 3 Rd Ree3 5 0.

where Kv, Kp [ R333 are positive definite matrices, (31)


a 5 3 a1 a2 a3 4 ^, where a1, a2, and a3 are distinct positive inte-
gers, and the map Writing R^ d Re 5 3 ri,j 4 i,j[ 51,2,36 [ SO 1 3 2 and expanding the

3 cross products, (31) is equivalent to a2r32 5 a3r23, a1r31 5


V a 1 R 2 ! a a i e i 3 1 R^
d Rei 2 , (23) a3r13, a1r21 5 a2r12. Since a1, a2, and a3 are positive, the rota-
i51
tion matrix R^ d Re can be written as
with 3 e1 e2 e3 4 the identity matrix. The feedback controller
(22) requires the full attitude R [ SO 1 3 2 and angular veloc- r11 r12 r13
ity v [ R3, and it is continuous on SO 1 3 2 3 R3. The feed- a2
back controller (22) can be interpreted as introducing a R^ r
d Re 5 E a1 12
r22 r23U . (32)
potential though the attitude-dependent term and dissipa- a3 a3
tion through the angular-velocity-dependent term. Knowl- a1 r13 a2 r23 r33
edge of the rigid-body inertia is not required by the feedback
controller. While the dissipation term and the attitude- Since R^
d Re is orthogonal it follows that
dependent term in the feedback controller appear linearly,
the feedback controller is nonlinear since the rotation matri-
a22 a23
ces are not elements of a vector space. a1 2 br212 1 a1 2 br213 5 0, (33)
a21 a21
The closed-loop full-attitude dynamics using the con-
troller (22) are a23 a23
a1 2 br213 1 a1 2 br223 5 0, (34)
a21 a22
#
Jv 5 Jv 3 v 2 Kvv 2 KpV a 1 R 2 , (24) a22 a23
# a1 2 br212 2 a1 2 br223 5 0. (35)
R 5 Rv3 . (25) a21 a22

JUNE 2011 « IEEE CONTROL SYSTEMS MAGAZINE 39


Since a1, a2, and a3 are distinct, the solution to (33)–(35) Let the perturbation in the initial attitude be given as
R 1 0, e 2 5 ReeeU0 , where R^
3
satisfies either a) r12 5 r13 5 r23 5 0 or b) none of the rij, i 2 j d Re [ S is given in (26) and
are zero. To see this result, first note that a) is a trivial U0 [ R3; see “Linearization of Nonlinear Models on Euclid-
solution of (33)–(35). Suppose r12 2 0. Then since a1, a2, ean Spaces and Non-Euclidean Spaces” for details. The per-
and a3 are distinct, it follows that 1 1 2 1 a22 /a21 22 2 0, turbation in the initial angular velocity is given as
1 1 2 1 a23 /a21 22 2 0, and 1 1 2 1 a23 /a22 22 2 0. Then, since r12 2 0, v 1 0, e 2 5 ev 0, where v 0 [ R3. Note that if e 5 0 then
(33) yields that r13 2 0, and from (34), it follows that 1 R 1 0, 0 2 , v 1 0, 0 22 5 1 Re, 0 2 , and hence
r23 2 0. Similar arguments hold for the case r13 2 0 and
r23 2 0. Thus, every solution to (33)–(35) satisfies either a) 1 R 1 t, 0 2 , v 1 t, 0 22 ; 1 Re, 0 2 (36)
or b).
Next, consider case b). Suppose a1 . a2. Then, for all time t [ 3 0, ` 2 , thereby obtaining the equilibrium
1 1 2 1 a22 /a21 22 r212 . 0 since r12 2 0, and hence (33) yields solution.
1 1 2 1 a23 /a21 22 r213 , 0. Since r213 is positive, it follows that Next, consider the solution to the perturbed equations of
a3 . a1. Thus, a3 . a1 . a2. This inequality implies motion for the closed-loop system (24)–(25). These equa-
tions are given by
a23 a23
a1 2 br213 , 0, a1 2 br223 , 0. #
2
a1 a22 Jv 1 t, e 2 5 Jv 1t, e 2 3 v 1 t, e 2 2 Kvv 1 t, e 2 2 KpV a 1R 1 t, e 22 , (37)
#
R 1 t, e 2 5 R 1 t, e 2 v 1 t, e 2 3 . (38)
Therefore, the left-hand side of (34) is negative, yielding a
contradiction. Similar arguments show that for a1 , a2, To linearize (24) – (25), we differentiate both sides of
(33)–(35) yields a contradiction for case b). Hence since a1, (37) and (38) with respect to e and substitute e 5 0,
a2, and a3 are distinct, case b) yields a contradiction, and yielding
the only solution to (33) – (35) is case a), that is,
#
r12 5 r13 5 r23 5 0. Jv e 1 t, 0 2 5 2Kv v e 1 t, 0 2 2 KpV a 1Re 1 t, 0 22 , (39)
Then, substituting r12 5 r13 5 r23 5 0 into (32), it follows #
Re 1 t, 0 2 5 Rev e 1 t, 0 2 3 , (40)
that r11, r22, r33 each have value 1 1 or 2 1. Thus, R^ d Re [ S
as given in (26), and hence 1 Re, 0 2 [ ~ in (27). u where v e 1 t, 0 2 ! 1'v 1t, e 2 / 1'e 22 e50 and Re 1 t, 0 2 ! 1'R 1 t, e 2 /
Proposition 1 shows that the continuous feedback con- 1'e 22 e50. Now we define the linearized states Dv, DU [ R3
troller results in a closed-loop system with four distinct as Dv 1 t 2 ! v e 1 t, 0 2 and DU 1 t 2 3 ! R^ e Re 1 t, 0 2 . Then from
equilibria. One of these equilibria is the desired equilib- (40) we obtain
rium specified by the attitude Rd, while the remaining # #
equilibria are designated by Ra, Rb, and Rg. These attitudes e Re 1t, 0 2 5 v e 1 t, 0 2 5 Dv 1 t 2 .
DU 1 t 2 3 5 R^ 3 3

differ from the desired attitude Rd by 180° of rotation about


each of the three body-fixed axes. Suppressing the time dependence, we obtain
#
Local Structure of the Closed-Loop System DU 5 Dv. (41)
We next analyze the stability of the desired equilibrium
and each of the three additional equilibria, and we deduce From (39), we obtain
the local structure of the closed-loop vector field near
#
each of these equilibria using the Lie group properties of JDv 5 2KvDv 2 KpV a 1 ReDU 3 2 . (42)
SO 1 3 2 . Since dim 3 SO 1 3 2 3 R3 4 5 6, each linearization
evolves on R6. To obtain the local structure of the closed- Now,
loop attitude dynamics, we linearize the closed-loop
equations about each equilibrium. While it is possible to V a 1 ReDU 3 2 5 a1e1 3 1 R^
d ReDU e1 2 1 a2e2 3 1 Rd ReDU e2 2
3 ^ 3

study the linearized dynamics in local coordinates, the 1 a 3 e 3 3 1 R^


d ReDU e3 2 ,
3

closed-loop vector field (24)–(25) can also be linearized


5 2 1 a1e13 R^ 3 3 ^ 3
d Ree 1 1 a 2 e 2 Rd Ree 2
using the Lie-group properties of the non-Euclidean
d Ree3 2 DU,
1 a3e33 R^ 3
manifold SO 1 3 2 . Let 1 Re, 0 2 [ ~ be an equilibrium solu- (43)
tion of the closed-loop system (24)–(25). Consider a per-
turbation of the equilibrium 1 Re, 0 2 [ ~ in terms of a where R^d Re belongs to S as given in (26) in Proposition 1.
perturbation parameter e [ R. We can express the per- Combining (41)–(43), we obtain the linearization of the
turbation as a rotation matrix in SO 1 3 2 using exponential closed-loop system (24)–(25) about each equilibrium in (27)
coordinates [34]–[37]. This representation guarantees as
that the perturbation is a rotation matrix for all values of #
$
the perturbation parameter. JDU 1 KvDU 1 KDU 5 0, (44)

40 IEEE CONTROL SYSTEMS MAGAZINE » JUNE 2011


Linearization of Nonlinear Models on Euclidean Spaces and Non-Euclidean Spaces
inearization of a nonlinear model is defined from the re- similar analysis can be used to linearize the reduced rotational
L sponses to an infinitesimal perturbation of initial conditions
from a given equilibrium. For rigid-body rotational dynamics,
dynamics of a rigid body.
In contrast to the nonlinear perturbation function (S1) for
we require that the perturbations in initial conditions lie on the rigid-body rotational dynamics, we can choose F as a linear
manifold SO 1 3 2 3 R3 so that the solution evolves on the mani- additive vector perturbation for models whose states evolve
fold SO 1 3 2 3 R3. on Euclidean spaces. For attitude dynamics, while a linear
Let F: SO 1 3 2 3 R3 3 R S SO 1 3 2 3 R3 be the mapping additive perturbation is used for the angular velocity, the ex-
such that F 1 Re, v e, e 2 represents the perturbation from the ponential function is used for the attitude perturbation in (S1).
equilibrium 1 Re, v e 2 , where e denotes the perturbation param- The reason behind this choice is that, while linear additive
eter. Thus e 5 0 implies no perturbation from the equilibrium perturbations guarantee that, for all values of the perturba-
solution. If 1 R 1 t, e 2 , v 1 t, e 2 2 represents the unique solution tion parameter e, the resulting angular velocity perturbations
corresponding to the perturbation e, then the linearization is lie in the Euclidean space of angular velocities, adding rota-
obtained by differentiating this perturbed solution with respect tion matrices does not guarantee that the resulting matrix is
to e, where the perturbation function is defined as also a rotation matrix. Since the exponential map (2) of SO 1 3 2
guarantees that the resulting matrix is a rotation matrix [34],
F 1 Re, v e, e 2 ! 1 Ree eU0 , v e 1 ev 0 2
3
(S1) [14], [36], [37], an exponential function is used to construct
the attitude perturbation as shown in (S1). Thus, the attitude
and 1 U0, v 0 2 [ R3 3 R3. This computation yields the lineariza- perturbation in (S1) lies in SO 1 3 2 for all values of the perturba-
tion of the rigid-body rotational dynamics at the equilibrium. A tion parameter e.

where the constant matrix K, depending on the equilib- exponential convergence. Since the feedback controller is
rium attitude, is defined as continuous, the closed-loop vector field has the additional
equilibria 1 Ra, 0 2 , 1 Rb, 0 2 , and 1 Rg, 0 2 .
Kp diag 1 a2 1 a3, a1 1 a3, a1 1 a2 2 , if Re 5 Rd, We next want to understand the global structure of the
2 Kpdiag 1 a2 1 a3, a1 2 a3, a1 2 a2 2 , if Re 5 Ra, solutions of the closed-loop system. Consider again the
K5 µ
2 Kpdiag 1 a3 2 a2, a3 2 a1, a1 1 a2 2 , if Re 5 Rb, closed-loop linearizations (44). For each linearization about
2 Kpdiag 1 a2 2 a3, a1 1 a3, a2 2 a1 2 , if Re 5 Rg. the unstable equilibria 1 Ra, 0 2 , 1 Rb, 0 2 , and 1 Rg, 0 2 , the linear-
(45) ization has at least one eigenvalue with a negative real part
Thus, the local attitude dynamics near each of the four equi- and at least one eigenvalue with a positive real part. Further-
librium solutions are given by a linear second-order more, the four equilibria are hyperbolic, that is, the lineariza-
mechanical system with positive damping. The stiffness tion has no imaginary eigenvalues. Thus, local properties of
depends on the equilibrium; near the desired equilibrium, the closed-loop vector field near each of these equilibria are
the stiffness is positive definite, and hence, the local attitude as follows. There exist stable and unstable manifolds for each
dynamics near the desired equilibrium are locally exponen- of the equilibria 1 Ra, 0 2 , 1 Rb, 0 2 , and 1 Rg, 0 2 , denoted by
tially stable. Around the additional three equilibria corre- 1 Was, Wau 2 , 1 Wbs, Wbu 2 , and 1 Wgs, Wgu 2 such that all solutions
sponding to the attitudes Ra, Rb, and Rg, the stiffness matrix that start on the stable manifold converge to the correspond-
has at least one negative eigenvalue resulting in a saddle ing unstable equilibrium, and all solutions that start on the
point. Therefore, the closed-loop dynamics near each equi- unstable manifold locally diverge from the corresponding
librium solution are unstable. Furthermore, these equilibria equilibrium [38]. Furthermore, the dimensions of the stable
have no zero eigenvalues, yielding the following result. and unstable manifolds of each unstable equilibrium are the
same as the number of eigenvalues with negative and posi-
Proposition 2 tive real parts, respectively [38].
The equilibrium 1 Rd, 0 2 of the closed-loop equations (24)– Since the union of the stable manifolds has dimension
(25) is asymptotically stable with locally exponential conver- less than six, it can be shown [23] that the set has measure
gence. The additional closed-loop equilibria 1 Ra, 0 2 , 1 Rb, 0 2 , zero and is nowhere dense. Furthermore, using Lyapunov
and 1 Rg, 0 2 are hyperbolic and unstable. analysis [23], it can be shown that all solutions converge to
one of the four equilibria. Since all solutions that converge
Global Analysis of the to the three unstable equilibria lie in a nowhere dense set,
Full-Attitude Closed-Loop System almost all closed-loop solutions converge to the desired
The feedback controller (22) renders the desired equi- equilibrium 1 Rd, 0 2 . This result is precisely stated in the fol-
librium 1 Rd, 0 2 locally asymptotically stable with local lowing theorem.

JUNE 2011 « IEEE CONTROL SYSTEMS MAGAZINE 41


Theorem 1 dent term and dissipation through the angular
Consider the closed-loop rigid-body attitude equations velocity-dependent term. Knowledge of the rigid-body
of motion given by (24) and (25) for the controller (22), inertia is not required by the feedback controller. Note
where Kp and Kv are positive definite, and a1, a2, and a3 that while the dissipation term and the reduced-attitude-
are distinct positive integers. Then the closed-loop equi- dependent term in the feedback controller appear lin-
librium 1 Rd, 0 2 is asymptotically stable with locally early, the feedback controller is nonlinear since the
exponential convergence. Furthermore, there exists a reduced-attitude vectors lie in S2 and, hence, are not ele-
nowhere dense set M such that all solutions starting in ments of a vector space.
M converge to one of the unstable equilibria The closed-loop reduced-attitude equations of motion
1 Ra, 0 2 , 1 Rb, 0 2 , and 1 Rg, 0 2 , and all remaining solutions are
converge to the desired equilibrium 1 Rd, 0 2 . Thus, the #
Jv 5 Jv 3 v 1 kp 1 Gd 3 G 2 2 Kvv, (49)
domain of attraction of the desired equilibrium 1 Rd, 0 2 is #
almost global. G 5 G 3 v. (50)
Since the domain of attraction of 1 Rd, 0 2 is the comple-
ment of the union of the stable manifolds for each of the Closed-Loop Equilibria
three unstable equilibria, the eigenspace corresponding Consider the equilibrium structure of the closed-loop
to the stable eigenvalues of an unstable equilibrium pro- equations (49) and (50). Equating the right-hand side of
vides explicit local information on the complement of the (49) and (50) to zero yields that an equilibrium at 1 Ge, v e 2
domain of attraction of 1 Rd, 0 2 near that unstable equilib- satisfies v e 5 mGe, where m is a constant, and
rium. Note that the structure and characteristic of the
unstable equilibria depends on the particular choice of m2JGe 3 Ge 1 kp 1 Gd 3 Ge 2 2 mKvGe 5 0.
feedback control. However, as shown in [9], the comple-
ment of the domain of attraction of 1 Rd, 0 2 is necessarily Premultiplying both sides by G^ ^
e yields mGe KvGe 5 0. Since
nonempty for every continuous time-invariant feedback Kv is positive definite and 7Ge 7 5 1, it follows that m 5 0.
controller on SO 1 3 2 3 R3. Thus, the existence of a nowhere Thus, v e 5 0. Then, we see that Ge 3 Gd 5 0, and hence
dense set, which is the complement of the domain of Ge 5 6Gd. We thus have the following proposition.
attraction of 1 Rd, 0 2 , is not a consequence of the specific
controller (22). Since global stabilization using continu- Proposition 3
ous time-invariant feedback is impossible, almost global The reduced closed-loop attitude equations of motion (49)
stabilization is the best possible result for this stabiliza- and (50) have two equilibrium solutions given by 1 Gd, 0 2
tion problem. and 1 2Gd, 0 2 .
Proposition 3 shows that the continuous time-invari-
REDUCED-ATTITUDE STABILIZATION ant feedback control results in a closed-loop system with
Rigid-body reduced-attitude stabilization is analyzed two distinct equilibria given by the desired equilibrium
using the fully actuated rigid-body equations of motion 1 Gd, 0 2 and an additional closed-loop equilibrium
1 2Gd, 0 2 . The existence of the two reduced-attitude equi-
#
Jv 5 Jv 3 v 1 u, (46) libria implies that there is a manifold of full-attitude
#
G 5 G 3 v, (47) equilibria, all of which satisfy R^b 5 Gd, and another dis-
joint manifold of full-attitude equilibria, all of which sat-
where G [ S2 and v [ R3. isfy R^b 5 2Gd; see “Relation Between Full Attitude and
The objective of reduced-attitude stabilization is to select a Reduced Attitude.”
continuous feedback controller u : S2 3 R3 S R3 that
asymptotically stabilizes the desired reduced-attitude Local Structure of the Closed-Loop System
equilibrium defined by the reduced-attitude vector Gd [ S2 We study linearizations of the closed-loop equations (49)
and angular velocity vector v 5 0. The feedback controller and (50) about the equilibrium 1 Gd, 0 2 and about the equi-
is given by librium 1 2Gd, 0 2 . Since dim 3 S2 3 R3 4 5 5, each lineariza-
tion evolves on R5. Similar to the analysis presented in [23],
u 5 kp 1 Gd 3 G 2 2 Kv v, (48) it can be shown that each equilibrium solution has the fol-
lowing local properties.
where kp is a positive real number and Kv is a positive
definite matrix. This feedback controller is based on Proposition 4
feedback of the reduced attitude G 5 R^b [ S2 and the The equilibrium 1 Gd, 0 2 of the closed-loop reduced-attitude
angular velocity v [ R3, and it is continuous on S2 3 R3. equations (49) and (50) is asymptotically stable and the con-
The feedback controller (48) can be interpreted as intro- vergence is locally exponential. Furthermore, the equilib-
ducing a potential through the reduced-attitude-depen- rium 1 2Gd, 0 2 is hyperbolic and unstable.

42 IEEE CONTROL SYSTEMS MAGAZINE » JUNE 2011


Global Analysis of the Reduced-Attitude
Closed-Loop System Relation Between Full Attitude
In this section, we study the global behavior of the closed- and Reduced Attitude
loop system defined by (49) and (50). The results are paral-
F or a rigid body, one reduced attitude corresponds to a
lel to those stated above for the full-attitude closed-loop manifold of full attitudes. To see this connection, note that
system. In particular, we characterize the domain of attrac- the definition of the reduced attitude G [ S2 implies that there
tion of the desired equilibrium 1 Gd, 0 2 and describe the exists a full attitude R0 [ SO 1 3 2 such that RT0 b 5 G. Then
global structure of the solutions that converge to the unsta- Rodrigues’s formula shows that every full attitude given by
ble equilibrium 1 2Gd, 0 2 . R 5 R0 exp 1 aG 3 2 with a [ R also satisfies RTb 5 G. In physi-
It can be shown using Lyapunov analysis that all cal terms, all full attitudes that differ by a rotation about G are
closed-loop solutions converge to one of the two closed- equivalent in the sense that they have the same reduced at-
loop equilibria (see “Lyapunov Analysis on SO 1 3 2 3 R3 titude, namely, G.
and S2 3 R3”). Since 1 2Gd, 0 2 is unstable and the linear-
ization about 1 2Gd, 0 2 has no imaginary or zero eigenval-
ues, it follows that there exists a stable invariant manifold
Mr [38] of the closed-loop equations (49) and (50) such of 1 Gd, 0 2 , is not a consequence of the specific controller
that all solutions that start in Mr converge to 1 2Gd, 0 2 . that is studied in this article. Rather, the complement of
The tangent space to this manifold at 1 2Gd, 0 2 is the the domain of attraction of 1 Gd, 0 2 is necessarily non-
stable eigenspace corresponding to the negative eigen- empty for every continuous time-invariant feedback con-
values. Since the linearization of the closed-loop dynam- troller on S2 3 R3 [9]. Thus, since global stabilization
ics at 1 2Gd, 0 2 has no eigenvalues on the imaginary axis, using continuous time-invariant feedback is impossible,
the closed-loop dynamics has no center manifold, and almost global stabilization is the best possible result for
every closed-loop solution that converges to 1 2Gd, 0 2 lies this stabilization problem.
in the stable manifold Mr.
It can be shown that Mr is a lower dimensional subset in CLOSED-LOOP PERFORMANCE LIMITATIONS
the five-dimensional manifold S2 3 R3 and is nowhere The difference between global asymptotic stability and
dense [23]. Hence, the closed-loop dynamics partition almost global asymptotic stability of the desired equilib-
S2 3 R3 into two sets. The first set Mr is defined by solu- rium of a closed-loop attitude control system leads to cer-
tions that converge to the unstable equilibrium 1 2Gd, 0 2 , tain closed-loop attitude performance limitations. Thus, we
and its complement is defined by the dense set of all solu- now address the third question posed in the section “Global
tions that converge to the desired equilibrium 1 Gd, 0 2 . Thus, Attitude Stabilization” related to performance implications
1 Gd, 0 2 is an asymptotically stable equilibrium of the closed- of almost global stabilization.
loop system (49) and (50) with an almost global domain of While all solutions that lie in the domain of attraction
attraction. asymptotically approach the desired equilibrium, solutions
that encounter a sufficiently small neighborhood of the
Theorem 2 stable manifold of one of the unstable hyperbolic equilibria
Consider the reduced rigid-body attitude dynamics can remain near that stable manifold for an arbitrarily long
given by the closed-loop equations (49) and (50). Assume
that kp is positive, and Kv is symmetric and positive defi-
nite. Then the closed-loop equilibrium 1 Gd, 0 2 is asymp- Lyapunov Analysis on SO 1 3 2 3 R3 and
totically stable with local exponential convergence. S2 3 R3

C
Furthermore, there exists a lower dimensional manifold losed-loop full-attitude dynamics evolve on the non-
Mr that is nowhere dense such that closed-loop solu- Euclidean manifold SO 1 3 2 3 R3, while closed-loop re-
tions starting in Mr converge to the unstable equilib- duced-attitude dynamics evolve on the non-Euclidean mani-
rium 1 2Gd, 0 2 and all remaining closed-loop solutions fold S2 3 R3. Since these manifolds are locally Euclidean,
converge to the desired equilibrium 1 Gd, 0 2 . Thus, the local stability properties of a closed-loop equilibrium solution
domain of attraction of the desired equilibrium is can be assessed using standard Lyapunov methods. In ad-
almost global. dition, the LaSalle invariance result and related Lyapunov
Since the domain of attraction of 1 Gd, 0 2 is the comple- results apply to closed-loop vector fields defined on these
ment of the stable manifold of the unstable equilibrium manifolds. However, since the manifolds SO 1 3 2 and S2 are
1 2Gd, 0 2 , the eigenspace corresponding to the stable compact, the radial unboundedness assumption cannot be
eigenvalues of 1 2Gd, 0 2 provides explicit local informa- satisfied; consequently, global asymptotic stability cannot fol-
tion on the complement of the domain of attraction of low from a Lyapunov analysis on Euclidean spaces [40], and
1 Gd, 0 2 near 1 2Gd, 0 2 . The presence of the nowhere dense therefore must be analyzed in alternative ways [19]–[23].
set, which is the complement of the domain of attraction

JUNE 2011 « IEEE CONTROL SYSTEMS MAGAZINE 43


loop properties of a continuous time-invariant vector field
3 on SO 1 3 2 3 R3 or S2 3 R3.
In contrast to continuous feedback controllers,
2.5 switched or discontinuous feedback controllers can also
be considered for attitude control. Controllers based on
2
nonsmooth feedback may not suffer from the perfor-
||θ|| (rad)

1.5 mance limitations observed in smooth controllers [27].


However, discontinuous controllers suffer from chatter-
1 ing at the surface of the discontinuity of the closed-loop
0.5 vector field. Chattering in discontinuous controllers can
be resolved by using hysteresis [39]. However, this dis-
0 continuity can render standard Lyapunov theory inappli-
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
t /T cable due to the non-Lipschitz nature of the closed-loop
(a) vector field [40], [41].
0.9
0.8 EXAMPLES OF RIGID-BODY ATTITUDE
0.7 MANEUVERS AND ATTITUDE STABILIZATION
0.6 We now illustrate the attitude control results by construct-
||ω|| (rad/s)

0.5 ing a rest-to-rest full-attitude maneuver, a rest-to-rest


0.4
reduced-attitude maneuver, a full-attitude stabilizing feed-
back controller, and a reduced-attitude stabilizing feedback
0.3
controller. The rigid body is assumed to be an asymmetric
0.2
body. The body-fixed frame is defined by the principal axes
0.1
of the body, and the body inertia matrix is assumed to be
0 J 5 diag 1 3, 4, 5 2 kg-m2.
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
t /T
(b) A Rest-to-Rest Full-Attitude Maneuver
2.4
We construct a rest-to-rest full-attitude maneuver that
2.2
aligns the three principal axes of the rigid body with the
2
three inertial axes. Specifically, the full-attitude maneuver
1.8
transfers the initial attitude R0 to a terminal attitude Rf,
||τ|| (N-m)

1.6
where
1.4
1.2
2 0.3995 0.8201 0.4097 1 0 0
1
0.8 R0 5 £ 0.1130 2 0.3995 0.9097 § , Rf 5 £ 0 1 0 § ,
0.6 0.9097 0.4097 0.0670 0 0 1
0.4
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 in a maneuver time of T 5 5 s. Equation (8) can be solved to
t /T
obtain a 5 1 0.5, 0.5, 0.707 2 , and the angle a 5 5p/6 rad.
(c)
Expressions for the attitude, angular velocity, and control
input that satisfy the specified boundary conditions are
FIGURE 1 (a) Principal angle, (b) norm of angular velocity, and given by (9)–(11). In this instance, the rotation angle is
(c) norm of the control torque plotted versus normalized time u 1 t 2 5 1 a/2 2 1 1 2 cos 11 p/T 2 t 22 rad.
for the rest-to-rest, full-attitude, open-loop maneuver for a
For the rest-to-rest full-attitude maneuver, Figure 1 shows
rigid body with principal moment of inertia matrix J 5 diag(3,
4, 5) kg-m 2 within the maneuver time T 5 5 s. The principal the time responses for the principal angle u, the norm of the
angle u(t) is obtained as a sinusoidal interpolation between angular velocity 7v 7, and the norm of the control torque 7u7
the given initial attitude R (0) and the desired final attitude versus normalized time t/T. Numerical solutions for these
R (T ) for this rest-to-rest maneuver. The angular velocity and open-loop maneuvers are obtained using Lie group varia-
control torque are obtained from the principal angle and
tional integrators, as described in “Numerical Integration of
resulting attitude profile.
Attitude Control Systems.” With this control scheme and
fixed initial and final attitudes, the normalized time plots for
period of time [38]. Such solutions in the domain of attrac- the principal angle, the angular velocity norm, and control
tion approach the desired equilibrium arbitrarily slowly. norm for every final time T qualitatively look similar to
This property of the closed-loop system is an attitude per- those shown in Figure 1, except that larger T requires less
formance limitation that arises due to the inherent closed- control effort.

44 IEEE CONTROL SYSTEMS MAGAZINE » JUNE 2011


Numerical Integration of Attitude Control Systems
quations (24) and (25) describe the full-attitude control sys- called variational integrators. In [S2]–[S4], Lie group variation-
E tem, whereas (49) and (50) describe the reduced-attitude
control system, which evolve on SO 1 3 2 3 R3 and on S2 3 R3,
al integrators are developed and used to numerically simulate
attitude dynamics and control systems. Lie group variational
respectively. Traditional numerical integration schemes, de- integrators incorporate the desirable features of both geometric
veloped for differential equations that evolve on Rn, do not integration schemes and variational integration schemes.
guarantee that the computed solution lies in the appropriate
manifold [S1]. In contrast to traditional numerical integration REFERENCES
[S1] E. Hairer, C. Lubich, and G. Wanner, Geometric Numerical Integra-
schemes, geometrical numerical integration schemes guaran- tion. New York: Springer-Verlag, 2000.
tee that the computed solution lies in the appropriate manifold [S2] T. Lee, M. Leok, and N. H. McClamroch, “Computational geometric
[S1]. Geometric integration schemes can be applied to mani- optimal control of rigid bodies,” Commun. Inform. Syst., vol. 8, no. 4. pp.
435–462, 2008.
folds such as S3 without requiring renormalization or projec- [S3] T. Lee, M. Leok, and N. H. McClamroch, “Lagrangian mechanics
tion. If the manifold under consideration is a Lie group such and variational integrators on two-spheres,” Int. J. Numerical Methods
as SO 1 3 2 , then such integration schemes are called Lie group Eng., vol. 79, pp. 1147–1174, 2009.
[S4] A . K. Sanyal and N. Nordkvist, “A robust estimator for almost glo-
methods. Integration methods that are obtained by using the bal attitude feedback tracking,” in Proc. AIAA Guidance, Navigation and
discrete version of the variational principle of mechanics are Control Conf., Toronto, Ontario, Aug. 2010, AIAA-2010-8344.

A Rest-to-Rest Reduced-Attitude Maneuver tity matrix. The feedback controller is given by (22), where
We construct a rest-to-rest reduced-attitude maneuver that we select the vector a 5 1 1, 2, 3 2 , and the nominal gains
aligns the third principal axis of the body frame with the Kp 5 diag 1 1, 2, 3 2 and Kv 5 diag 1 5, 10, 15 2 .
third axis of the inertial frame. This reduced-attitude The closed-loop eigenvalues of the linearized closed-
maneuver transfers the initial reduced attitude G0 5 loop system at the equilibrium 1 Rd, 0 2 are given
R 1 0 2 ^e3 5 1 0.9097, 0.4097, 0.0670 2 to the terminal reduced by 5214.3739,29.1231, 23.618, 21.382, 20.8769, 20.62616 ,
attitude Gf 5 1 0, 0, 1 2 in a maneuver time of T 5 5 s. The thereby describing the local convergence rate near the
axis and angle variables are given as a 5 1 0.9118, desired equilibrium. The system has three additional
2 0.4107, 0 2 and a 5 1.5038 rad, respectively. Expressions equilibria corresponding to the attitudes 5 diag 1 1, 21, 21 2 ,
for the reduced attitude, the angular velocity, and the con- diag 1 21,21, 1 2 , diag 1 21, 1, 21 26 as described by (28)–(30)
trol input that satisfy the specified boundary conditions in Proposition 1.
are given by (15)–(17). In this instance, the rotation angle is We refer to the gains given above as defining a nominal
u 1 t 2 5 1 a/2 2 1 1 2 cos 11 p/T 2 t 22 rad. controller. Two alternative sets of gains define a stiff con-
The full-attitude maneuver studied above also satis- troller and a damped controller. The stiff gains are obtained
fies the boundary conditions imposed for the reduced- by increasing Kp and decreasing Kv by 25% each. In a simi-
attitude maneuver for the specified initial full rigid-body lar manner, the damped gains are obtained by decreasing
attitude. Since the reduced-attitude maneuver is less con- Kp and increasing Kv by 25% each. In each case, the desired
strained than the full-attitude maneuver, it typically full-attitude equilibrium 1 Rd, 0 2 is almost globally asymp-
requires less control effort. For this rest-to-rest reduced- totically stable on SO 1 3 2 3 R3.
attitude maneuver, Figure 2 gives the (a) time responses The three sets of gains yield nominal, stiff, and damped
of the principal angle, (b) norm of the angular velocity, closed-loop responses locally near the desired equilibrium.
and (c) norm of the control torque, all plotted versus nor- These closed-loop responses are shown in Figure 4 for the
malized time. Figure 2(d) shows the reduced-attitude initial conditions
vector components versus normalized time for this
maneuver. Figure 3 gives three snapshots of the reduced- 0 0.2887 0.4082 2 0.8660
attitude configuration, and the path traced out on the v 1 0 2 5 £ 0 § , R 1 0 2 5 £ 2 0.8165 0.5774 0 § . (51)
two-sphere at different instants in this maneuver, includ- 0 0.5000 0.7071 0.5000
ing the initial and final time instants. In Figure 3, the axes
shown in the snapshots of the reduced-attitude maneu- The error in the attitude is represented by the eigenangle
vers are the body axes.
1
U 1 t 2 5 cos21 1 trace 1 R 1 t 22 2 12 , (52)
2
Full-Attitude Stabilization
We construct a feedback controller that asymptotically which is a scalar measure of the error between R and I.
aligns the three principal axes of the rigid body with the Thus, U 1 0 2 5 80°. From Figure 4(a), we see that, com-
three inertial axes. Thus the desired attitude Rd is the iden- pared to the nominal controller, the stiff controller acts

JUNE 2011 « IEEE CONTROL SYSTEMS MAGAZINE 45


1.6 0.5
1.4 0.45
1.2 0.4
0.35

||ω|| (rad/s)
||θ|| (rad) 1 0.3
0.8 0.25
0.6 0.2
0.15
0.4
0.1
0.2 0.05
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
t /T t /T
(a) (b)
1.4 1.2
1.2 1

Γ (Unit Vector)
1 0.8 Γ1
||τ|| (N-m)

0.8 0.6 Γ2
Γ3
0.6 0.4
0.4 0.2
0.2 0
0 −0.2
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
t /T t /T
(c) (d)

FIGURE 2 (a) Principal angle, (b) norm of angular velocity, (c) norm of the control torque, and (d) components of the reduced-attitude
vector versus normalized time for the rest-to-rest reduced-attitude open-loop maneuver for a rigid body with inertia matrix J 5 diag(3,
4, 5) kg-m2, within the maneuver time T 5 5 s. The reduced-attitude vector is the direction of the third principal axis of the rigid body,
and the maneuver aligns the reduced-attitude vector with the third coordinate axis of the inertial coordinate frame. The principal axis of
rotation is the cross product of the given initial reduced attitude G(0) and the desired final reduced attitude G(T), while the principal angle
u(t) is obtained by a sinusoidal interpolation between the boundary conditions. The angular-velocity and control-torque profiles are
obtained from the principal angle and the resulting reduced-attitude profile. Although the norms of the angular-velocity and control-
torque profiles look similar in shape to the corresponding profiles for the full-attitude maneuver, their values are considerably lower.

faster in reducing the error to zero, while the damped to the stable and unstable manifolds at the equilib-
controller is slower in its response. However, as shown rium 1 Rg, 0 2 .
in Figure 4(b) and (c), the stiff controller requires a To demonstrate performance limitations, we choose a
higher control torque and also leads to larger transient family of initial conditions that lie close to the stable mani-
angular velocities compared to the nominal controller, fold of the unstable equilibrium 1 Rg, 0 2 . We choose
while the damped controller requires less control torque va 5 3 DUa Dv a 4 and vb 5 3 DUb Dv b 4 as a linear combination
and has lower transient angular velocities. The param- of eigenvectors from the stable and unstable eigenspaces,
eters a1, a2, and a3 distribute the control effort among the respectively, of the equilibrium 1 Rg, 0 2 , where
three body axes, and these gains can be modified
accordingly based on the moment of inertia and the 0.4021 0.5939
desired response. DUa 5 £ 0.3037 § , Dv a 5 £ 2 0.9528 § ,
To illustrate the closed-loop performance limitations, 0.2993 2 0.9542
we consider the linearization of the closed-loop system, as
0 0
given in (44) for the nominal gains, at the unstable equilib-
DUb 5 £ 0.8432 § , Dv b 5 £ 0.5375 § .
rium corresponding to the attitude Rg 5 diag 1 21, 1, 21 2 .
0.9827 0.1849
The local closed-loop dynamics near this unstable
equ i librium are described by the positive eigen- Consider the initial conditions
v a l u e s 1 0.1882, 0.6375 2 , and negative eigenvalues
120.2324, 21.4343,23.1882, 23.1375 2 . These eigenvalues R 1 0 2 5 diag 1 21, 1, 212 exp 1 dDUa3 1 eDUb3 2 , (53)
confirm that the equilibrium 1 Rg, 0 2 is unstable, and the v 1 0 2 5 dDv a 1 eDv b, (54)
stable manifold has dimension four, while the unstable
manifold is two dimensional. The eigenspaces correspond- which depends on the parameter e. If we choose 0 , d V 1,
ing to the negative and positive eigenvalues are tangent then, for e 5 0, Figure 5 shows that the initial conditions

46 IEEE CONTROL SYSTEMS MAGAZINE » JUNE 2011


t = 0.0 80
Nominal
Damped
60 Stiff
1
0.6

Θ (°)
40
0.2
z

–0.2
–0.6 20
–1
0
0 2 4 6 8
–1

–1
1 –0.6
6 Time (s)
–0.6
6 –0.2
2
–0.2 0 0.2 (a)
0.2
20
y 0.6
61 0.6
10 6 x
(a)

ω 1 (°/s)
0
t = 2.5 −20
−40
0 2 4 6 8
1 40

ω 2 (°/s)
Nominal
0.6 20 Damped
0.2 0 Stiff
z

–0.2 0 2 4 6 8
–0.6 30
ω 3 (°/s)

–1 20
10
0

–1
1 –1
– 0 2 4 6 8
–0.6
6–0.2 –0.6
6 Time (s)
–0.2
2
0.2 0.6 0.2
y 1 1 0.6 x (b)
(b) 10
Nominal
t = 5.0
Two−Norm of u (N-m)

8 Damped
Stiff
6
1
0.6 4
0.2
z

–0.2 2
–0.6
0
–1 0 2 4 6 8
Time (s)

–1 (c)
–1
1 –0.6
6
–0.6
6–0.2 –0.2
2
0.2 0 0.2
y 61
0.6 10 6
0.6 x
FIGURE 4 (a) Eigenangle, (b) angular velocity components along
(c) each body-fixed axis, and (c) the two-norm of the control torque
are plotted for three controllers, namely, a nominal controller, a stiff
FIGURE 3 Snapshots of the rest-to-rest, open-loop, reduced-atti- controller, and a damped controller. As shown in (a), the error con-
tude maneuver for a rigid body with inertia matrix J 5 diag(3, 4, 5) verges to zero faster in the case of the stiff controller since the gain
kg-m2. The reduced-attitude vector is denoted by a bold green line Kp that acts on the error is 25% larger than the nominal controller.
from the origin to a point on the surface of the two-sphere. The On the other hand, the error converges more slowly for the damped
path traced by the tip of the reduced-attitude vector on the two- controller where the gain Kp is 25% lower than the nominal case.
sphere at several instants during the maneuver time T 5 5 s is While the stiff controller yields faster convergence, the angular
given by a bold red line. The principal axis of rotation, which is velocity as shown in (b) has larger transients in the case of the stiff
inertially fixed, is given by the blue line. The indicated axes are the controller compared to the nominal or damped controller. Further-
body axes and not the inertial axes. (a) shows the initial reduced more, as shown in (c), the magnitude of the torque for the stiff
attitude, (b) is the reduced attitude at time t 5 2.5 s, and (c) is the controller is higher than the nominal controller, while the magni-
final reduced attitude at time t 5 5 s. tude of the control torque is lower for the damped controller.

JUNE 2011 « IEEE CONTROL SYSTEMS MAGAZINE 47


if va is not zero, then, whether or not e 5 0, the initial con-
2D-Unstable ditions (53), (54) lie close to the stable manifold but never
Manifold exactly at the equilibrium 1 Rg, 0 2 . In the simulations, we
vb choose d 5 10 24.
vb In Figure 6, we plot the scalar attitude error given by the
eigenangle (52) for four initial conditions corresponding to
ε 4D-Stable
Manifold four different values of e in (53) and (54) given by
va
δ e [ 5 1025, 1024, 1023, 1022 6 . As shown in Figure 6, the time
of convergence increases with decreasing values of e. Fur-
(Rγ , 0)
thermore, the increase in the transient period occurs when
the attitude error is near 180°, indicating that the solutions
remain close to the unstable equilibrium 1 Rg, 0 2 for an
extended period. This behavior demonstrates that the
FIGURE 5 Initial conditions to illustrate closed-loop performance
closer the solution starts to the complement of the domain
limitations. The unstable equilibrium (Rg, 0) has a four-dimen-
sional stable manifold and a two-dimensional unstable manifold. of attraction of the desired equilibrium 1 Rd, 0 2 , the longer it
The vectors va and vb are tangent to the stable and the unstable takes for the solution to converge to 1 Rd, 0 2 .
manifolds, respectively, at (Rg, 0). The initial conditions are pertur-
bations from (Rg, 0) such that, for e 5 0, the initial condition lies Reduced-Attitude Stabilization
close to the stable manifold, and, as e increases, the initial condi-
We construct a feedback controller that asymptotically
tion lies farther away from the stable manifold. A fixed nonzero d
guarantees that the initial condition is near but not exactly at the aligns the third principal axis of the body frame with the
equilibrium (Rg, 0) for every sufficiently small e. third axis of the inertial frame. Thus, b 5 1 0, 0, 1 2 , and the
goal is to asymptotically stabilize Gd 5 1 0, 0, 1 2 , that is, to
asymptotically align G 5 R^b with Gd. The feedback con-
troller, given by (48), is chosen to have the gains
180
Kv 5 diag 1 5, 10, 15 2 and kp 5 4.
160 ε = 1e−5
140 ε = 1e−4 The closed-loop eigenvalues of the linearized closed-
ε = 1e−3 loop system at 1 Gd, 0 2 are given by 520.8333 6 0.7993i,
120 ε = 1e−2
100 20.5, 22, 23 6 , thereby describing the local conver-
Θ (°)

80 gence rate near the desired equilibrium. There also


60 exists one additional equilibrium given by 1 2Gd, 0 2
40 that is unstable.
20 We refer to the gains given above as defining a nominal
0 controller. Two alternative sets of gains define a stiff con-
0 5 10 15 20 25 30 35 40
Time (s) troller and a damped controller. The stiff gains are obtained
by increasing kp and decreasing Kv by 25% each. In a similar
FIGURE 6 The error U(t) for the nominal controller for four initial manner, the damped gains are obtained by decreasing kp
conditions corresponding to four different values of e in (53) and and increasing Kv by 25% each. In each case, the desired
(54). The time for the error to converge to zero increases as e is reduced-attitude equilibrium 1 Gd, 0 2 is almost globally
decreased. Most of the increase in the transient time occurs in the asymptotically stable on S2 3 R3.
early part of the trajectory, where the scalar error is close to 180°.
The three sets of gains yield nominal, stiff, and damped
This behavior indicates that, as e decreases, the trajectory dwells
close to the unstable equilibrium (Rg, 0) and hence close to the closed-loop responses locally near the desired equilib-
error of 180°, since the solution starts closer to the four-dimen- rium. Closed-loop responses are plotted in Figure 7 for the
sional stable manifold as illustrated in Figure 5. However, since same initial conditions as above for the full-attitude case.
the initial condition does not lie on the stable manifold, it soon Thus the initial condition for the reduced attitude is
diverges along the two-dimensional unstable manifold, finally con-
G 1 0 2 5 R 1 0 2 ^e3 5 1 0.5, 0.7071, 0.5 2 . The error in the reduced
verging to the desired equilibrium (Rd , 0). The divergence along
the unstable manifold results in a sharp drop in the attitude error attitude is
to zero as shown above. The smaller the parameter e, the longer
the trajectories dwell close to the unstable equilibrium (Rg, 0). U 1 t 2 5 cos21 1 G^
d G 1 t 22 , (55)

which is a scalar measure of the error between G and Gd rep-


(53), (54) lie close to the stable manifold of the unstable resenting the angle between the two unit vectors. Thus,
equilibrium. As e increases, the distance to the stable U 1 0 2 5 60°. From Figure 7(a), we see that, compared to the
manifold of the unstable equilibrium 1 Rg, 0 2 increases. nominal controller, the stiff controller is faster in reducing
Therefore, as e decreases, we expect the time of conver- the error to zero, while the damped controller is slower in its
gence to the desired attitude Rd 5 I to increase. Note that response. However, as shown in Figure 7(b) and (c), the stiff

48 IEEE CONTROL SYSTEMS MAGAZINE » JUNE 2011


controller requires a higher control torque and also leads to
larger transient angular velocities compared to the nominal 60
controller, while the damped controller requires less control Nominal
Damped
torque and has lower transient angular velocities. Stiff
To illustrate performance limitations in reduced-atti- 40
tude feedback, we consider the linearization of the closed-

Θ (°)
loop system for the nominal controller about the unstable
equilibrium 1 2Gd, 0 2 . It can be shown that the linearized 20
closed-loop dynamics are described by the positive eigen-
values 1 0.3508, 0.5907 2 and the negative eigenvalues
1 22.2573, 22.8508, 23 2 . These eigenvalues indicate that 0
the equilibrium at 1 2Gd, 0 2 is unstable, and the stable man- 0 5 10
Time (s)
ifold has dimension three, while the unstable manifold is
(a)
two dimensional. The eigenspaces corresponding to the
negative and positive eigenvalues are tangent to the stable

ω 1 (°/s)
0
and unstable manifolds at the equilibrium 1 2Gd, 0 2 . −20
To demonstrate performance limitations, we choose a −40
family of initial conditions that lie close to the stable mani- 0 2 4 6 8 10
fold of the unstable equilibrium 1 2Gd, 0 2 . We choose 20

ω 2 (°/s)
10
va 5 3 DUa Dv a 4 and vb 5 3 DUb Dv b 4 as a linear combination
0
of eigenvectors from the stable and unstable eigenspaces, −10
respectively, of the equilibrium 1 2Gd, 0 2 , where 0 2 4 6 8 10
1
ω 3 (°/s)
Nominal
0 0 0 Damped
Stiff
DUa 5 £ 0 § , Dv a 5 £ 0 § , −1
2 0.3162 0.9487 0 2 4 6 8 10
Time (s)
0.8610 0.5086 (b)
DUb 5 £ 0 § , Dv b 5 £ 0 § .
10
0 0 Nominal
Two−Norm of u (N-m)

8 Damped
Stiff
Consider the family of initial conditions
6
G 1 0 2 5 exp 1 2dDUa3 2 eDUb3 2 1 2Gd 2 , (56)
4
v 1 0 2 5 dDv a 1 eDv b, (57)
2
which depend on the parameter e.
If we choose 0 , d V 1, then, for e 5 0, the initial condi- 0
tions (56), (57) lie close to the stable manifold of the unsta- 0 5 10
ble equilibrium 1 2Gd, 0 2 . As e increases, the distance to the Time (s)
(c)
stable manifold of the unstable equilibrium 1 2Gd, 0 2
increases. Therefore, as e decreases, we expect the time of
FIGURE 7 (a) The angle between the desired and actual reduced-
convergence to the desired attitude Gd to increase. Note attitude vectors, (b) angular velocity components along each
that, if va is not zero, then, whether or not e 5 0, the initial body-fixed axis, and (c) the two-norm of the control torque plotted
conditions (56), (57) lie close to the stable manifold but for three controllers, namely, a nominal controller, a stiff controller,
never exactly at the equilibrium 1 2Gd, 0 2 . In the simula- and a damped controller. As shown in (a), the error converges to
zero faster in the case of a stiff controller since the gain kp that acts
tions, we choose d 5 10 24.
on the error is 25% larger than the nominal controller. On the other
In Figure 8, we plot the scalar reduced-attitude error hand, the error converges more slowly for the damped controller,
defined in (55) for four initial conditions corresponding to where the gain kp is 25% lower than the nominal case. While the
values of e in (56) and (57) given by e [ 5 1025, 1024, stiff controller yields faster convergence, the angular velocity as
1023, 1022 6 . As shown in Figure 8, the time of convergence shown in (b) has larger transients in the case of the stiff controller
compared to the nominal and damped controllers and hence can
increases as e decreases. Furthermore, the increase in the
excite unmodeled dynamics. Furthermore, as shown in (c), the
transient period occurs when the reduced-attitude error is magnitude of the torque for the stiff controller is higher than the
near 180°, indicating that the solutions remain close to the nominal controller, while the magnitude of the control torque is
unstable equilibrium 1 2Gd, 0 2 for an extended period. This lower for the damped controller.

JUNE 2011 « IEEE CONTROL SYSTEMS MAGAZINE 49


the domain of attraction lies in a lower dimensional set of
180 the state space that is nowhere dense. Some features of the
ε = 1e−5
160 ε = 1e−4 geometry of the complement of the domain of attraction of
140 ε = 1e−3
ε = 1e−2 the desired equilibrium are described, identifying it as the
120 union of the undesired closed-loop equilibria and their
100
Θ (°)

associated stable manifolds. A key aspect of the analysis


80
given in this article is that we perform both local and global
60
40 analysis of the feedback controllers without resorting to
20 coordinates for SO 1 3 2 or S2.
0
0 5 10 15 20 25 30 35 40 ACKNOWLEDGMENTS
Time (s)
We appreciate the many conversations and suggestions
made by Taeyoung Lee and Melvin Leok, who helped to
FIGURE 8 The error U(t) for the nominal controller for four initial clarify some of the attitude control issues addressed in this
conditions corresponding to four different values of e in (56) and
(57). The time for the error to converge to zero increases as e is
article. We also thank the reviewers for suggestions that
decreased. Most of this increase in the transient time occurs in the helped to improve the presentation of this article.
early part of the trajectory, where the scalar error is close to 180°.
This behavior indicates that, as e decreases, the trajectory dwells AUTHOR INFORMATION
close to the unstable equilibrium (2Gd, 0), and hence close to the Nalin A. Chaturvedi (nalin.chaturvedi@us.bosch.com) is
error of 180°, since the solution starts closer to the three-dimen-
sional stable manifold. However, since the initial condition does
a senior research engineer at the Research and Technol-
not lie on the stable manifold, it soon diverges along the two- ogy Center of the Robert Bosch LLC, Palo Alto, California.
dimensional unstable manifold, finally converging to the desired He received the B.Tech. and the M.Tech. in aerospace engi-
reduced-attitude equilibrium (Gd, 0). The divergence along the neering from the Indian Institute of Technology, Bombay,
unstable manifold results in a sharp drop in the attitude error to in 2003, receiving the Institute Silver Medal. He received
zero as shown above. The smaller the parameter e, the longer the
trajectories dwell close to the unstable equilibrium manifold.
the M.S. in mathematics and the Ph.D. in aerospace engi-
neering from the University of Michigan, Ann Arbor, in
example illustrates that the closer the solution starts to the 2007. He was awarded the Ivor K. McIvor Award in recog-
complement of the domain of attraction, the longer it takes nition of his research in applied mechanics. His current
to converge to 1 Gd, 0 2 . interests include the development of model-based control
for complex physical systems involving thermal-chemical-
CONCLUSIONS fluid interactions, nonlinear dynamical systems, state and
This article summarizes global results on attitude control parameter estimation, and adaptive control with applica-
and stabilization for a rigid body using continuous time- tions to systems governed by PDE, nonlinear stability the-
invariant feedback. The analysis uses methods of geomet- ory, geometric mechanics, and nonlinear control. He is a
ric mechanics based on the geometry of the special member of the Energy Systems subcommittee within the
orthogonal group SO 1 3 2 and the two-sphere S2. Mechatronics technical committee of the ASME Dynamic
We construct control laws for full- and reduced-atti- Systems and Controls Division and an associate editor on
tude maneuvers for arbitrary given boundary condi- the IEEE Control Systems Society Editorial Board. He is a
tions and an arbitrary maneuver time. These results visiting scientist at the University of California, San Diego.
demonstrate controllability properties for both the full- He can be contacted at Robert Bosch LLC, Research and
and for reduced-attitude models, assuming full actua- Technology Center, 4005 Miranda Avenue, Suite 200, Palo
tion. The open-loop controllers have a discontinuous Alto, CA 94304 USA.
dependence on the boundary conditions, thus suggest- Amit K. Sanyal received the B.Tech. in aerospace engi-
ing the impossibility of global stabilization using con- neering from the Indian Institute of Technology, Kanpur,
tinuous feedback. in 1999. He received the M.S. in aerospace engineering
Both full- and reduced-attitude feedback stabilization from Texas A&M University in 2001, where he obtained
results are provided, assuming full actuation. Since no con- the Distinguished Student Masters Research Award. He
tinuous time-invariant feedback controller can achieve received the M.S. in mathematics and Ph.D. in aerospace
global asymptotic stabilization of an equilibrium, global engineering from the University of Michigan in 2004.
properties of the closed-loop attitude dynamics using the From 2004 to 2006, he was a postdoctoral research asso-
continuous feedback controllers (22) and (48) are character- ciate in the Mechanical and Aerospace Engineering De-
ized. The particular feedback controllers given here are partment at Arizona State University. From 2007 to 2010,
globally defined and continuous, and they achieve almost- he was an assistant professor in mechanical engineering
global full-attitude stabilization or almost-global reduced- at the University of Hawaii and is currently an assistant
attitude stabilization in the sense that the complement of professor in mechanical and aerospace engineering at

50 IEEE CONTROL SYSTEMS MAGAZINE » JUNE 2011


New Mexico State University. His research interests are ics and Control of Multibody Systems, J. E. Marsden, P. S. Krishnaprasad, and J.
in geometric mechanics, geometric control, discrete vari- C. Simo, Eds. Providence, RI: American Mathematical Society.
[17] S. Bharadwaj, M. Osipchuk, K. D. Mease, and F. Park, “A geometric ap-
ational mechanics for numerical integration of mechani- proach to global attitude stabilization,” in Proc. AIAA/AAS Astrodynamics
cal systems, optimal control and estimation, geometric/ Specialist Conf., San Diego, 1996, pp. 806–816.
algebraic methods applied to nonlinear systems, space- [18] S. Bharadwaj, M. Osipchuk, K. D. Mease, and F. C. Park, “Geometry and
inverse optimality of global attitude stabilization,” J. Guid. Control Dyn.,
craft guidance and control, and control of unmanned vol. 21, no. 6, pp. 930–939, 1998.
vehicles. He is a member of the IEEE Control Systems [19] N. A. Chaturvedi, F. Bacconi, A. K. Sanyal, D. S. Bernstein, and N. H.
Society Technical Committee on Aerospace Control and McClamroch, “Stabilization of a 3D rigid pendulum,” in Proc. American
Control Conf., Portland, OR, 2005, pp. 3030–3035.
the AIAA Guidance, Navigation, and Control Technical [20] N. A. Chaturvedi and N. H. McClamroch, “Global stabilization
Committee. of an inverted 3D pendulum including control saturation effects,” in
N. Harris McClamroch received the Ph.D. in engineer- Proc. 45th IEEE Conf. Decision and Control, San Diego, CA, Dec. 2006,
pp. 6488–6493.
ing mechanics from the University of Texas at Austin. [21] N. A. Chaturvedi, N. H. McClamroch, and D. S. Bernstein, “Stabiliza-
Since 1967 he has been with the University of Michigan, tion of a specified equilibrium in the inverted equilibrium manifold of the
Ann Arbor, Michigan, where he is a professor emeritus 3D pendulum,” in Proc. American Control Conf., 2007, pp. 2485–2490.
[22] N. A. Chaturvedi and N. H. McClamroch, “Asymptotic stabilization
in the Department of Aerospace Engineering. During the of the hanging equilibrium manifold of the 3D pendulum,” Int. J. Robust
past 25 years, his primary research interest has been in Nonlinear Control, vol. 17, no. 16, pp. 1435–1454, 2007.
nonlinear control. He has worked on control engineering [23] N. A. Chaturvedi, N. H. McClamroch, and D. S. Bernstein, “Asymptotic
smooth stabilization of the inverted 3D pendulum,” IEEE Trans. Automat.
problems arising in robotics, automated systems, and aero- Contr., vol. 54, no. 6, pp. 1204–1215, 2009.
space flight systems. He is a Fellow of the IEEE, received [24] W. Kang, “Nonlinear H-infinity control and its application to rigid
the IEEE Control Systems Society Distinguished Member spacecraft,” IEEE Trans. Automat. Contr., vol. 40, no. 7, pp. 1281–1285, 1995.
[25] A. K. Sanyal and N. A. Chaturvedi, “Almost global robust attitude
Award, and is a recipient of the IEEE Third Millennium tracking control of spacecraft in gravity,” in Proc. AIAA Conf. Guidance,
Medal. He has served as an associate editor and editor of Navigation, and Control, Honolulu, HI, Aug. 2008, AIAA-2008-6979.
IEEE Transactions on Automatic Control, and he has held [26] A. K. Sanyal, A. Fosbury, N. A. Chaturvedi, and D. S. Bernstein, “Iner-
tia-free spacecraft attitude trajectory tracking with disturbance rejection
numerous positions in the IEEE Control Systems Society, and almost global stabilization,” AIAA J. Guid. Control Dyn., vol. 32, pp.
including president. 1167–1178, 2009.
[27] N. A. Chaturvedi and N. H. McClamroch, “Asymptotic stabilization of
the inverted equilibrium manifold of the 3D pendulum using non-smooth
feedback,” IEEE Trans. Automat. Contr., vol. 54, no. 11, pp. 2658–2662, 2009.
REFERENCES [28] J. L. Junkins and J. D. Turner, Optimal Spacecraft Rotational Maneuvers.
[1] M. D. Shuster, “A survey of attitude representations,” J. Astronaut. Sci., New York: Elsevier, 1986.
vol. 41, no. 4, pp. 439–517, 1993. [29] M. J. Sidi, Spacecraft Dynamics and Control—A Practical Engineering Ap-
[2] J. Stuelpnagel, “On the parameterization of the three-dimensional rota- proach. Cambridge, U.K.: Cambridge Univ. Press, 2000.
tion group,” SIAM Rev., vol. 6, no. 4, pp. 422–430, 1964. [30] J. Shen, A. K. Sanyal, N. A. Chaturvedi, D. S. Bernstein, and N. H. Mc-
[3] R. E. Mortensen, “A globally stable linear attitude regulator,” Int. J. Con- Clamroch, “Dynamics and control of a 3D pendulum,” in Proc. IEEE Conf.
trol, vol. 8, no. 3, pp. 297–302, 1968. Decision and Control, Bahamas, 2004, pp. 323–328.
[4] B. Wie and P. M. Barba, “Quaternion feedback for spacecraft large angle [31] N. A. Chaturvedi, T. Y. Lee, M. Leok, and N. H. McClamroch, “Non-
maneuvers,” AIAA J. Guid. Control Dyn., vol. 8, no. 3, pp. 360–365, 1985. linear dynamics of the 3D pendulum,” J. Nonlinear Sci., vol. 21, no. 1, pp.
[5] B. Wie, H. Weiss, and A. Arapostathis, “Quaternion feedback regulator for 3–32, 2011.
spacecraft eigenaxis rotation,” J. Guid. Control Dyn., vol. 12, pp. 375–380, 1989. [32] N. A. Chaturvedi, N. H. McClamroch, and D. S. Bernstein, “Stabiliza-
[6] J. T. Wen and K. K. Delgado, “The attitude control problem,” IEEE Trans. tion of a 3D axially symmetric pendulum,” Automatica, vol. 44, no. 9, pp.
Automat. Cont., vol. 36, no. 10, pp. 1148–1162, 1991. 2258–2265, 2008.
[7] R. Bach and R. Paielli, “Linearization of attitude-control error dynam- [33] P. Tsiotras, M. Corless, and J. Longuski, “A novel approach for the atti-
ics,” IEEE Trans. Automat. Cont., vol. 38, no. 10, pp. 1521–1525, 1993. tude control of an axisymmetric spacecraft subject to two control torques,”
[8] S. M. Joshi, A. G. Kelkar, and J. T.-Y. Wen, “Robust attitude stabilization Automatica, vol. 31, no. 8, pp. 1099–1112, 1995.
of spacecraft using nonlinear quaternion feedback,” IEEE Trans. Automat. [34] A. M. Bloch, Nonholonomic Mechanics and Control. New York: Springer-
Cont., vol. 40, no. 10, pp. 1800–1803, 1995. Verlag, 2003.
[9] S. P. Bhat and D. S. Bernstein, “A topological obstruction to continuous [35] V. S. Varadarajan, Lie Groups, Lie Algebras, and Their Representations.
global stabilization of rotational motion and the unwinding phenomenon,” New York: Springer-Verlag, 1984.
Syst. Control Lett., vol. 39, no. 1, pp. 63–70, 2000. [36] R. M. Murray, Z. Li, and S. S. Sastry, A Mathematical Introduction to Ro-
[10] P. C. Hughes, Spacecraft Attitude Dynamics. New York: Wiley, 1986. botic Manipulation. Boca Raton, FL: CRC Press, 1994.
[11] B. Wie, Spacecraft Vehicle Dynamics and Control. Reston, VA: AIAA, 1998. [37] F. Bullo and A. D. Lewis, Geometric Control of Mechanical Systems. New
[12] G. Meyer, “Design and global analysis of spacecraft attitude control York: Springer-Verlag, 2005.
systems,” NASA Tech. Rep. R-361, Mar. 1971. [38] J. Guckenheimer and P. Holmes, Nonlinear Oscillations, Dynamical Sys-
[13] P. E. Crouch, “Spacecraft attitude control and stabilization: Applica- tems, and Bifurcations of Vector Fields. New York: Springer-Verlag, 1983.
tions of geometric control theory to rigid body models,” IEEE Trans. Au- [39] C. G. Mayhew, R. G. Sanfelice, and A. R. Teel, “Robust global asymp-
tomat. Contr., vol. 29, no. 4, pp. 321–331, 1984. totic attitude stabilization of a rigid body by quaternion-based hybrid feed-
[14] F. Bullo, R. M. Murray, and A. Sarti, “Control on the sphere and reduced back,” in Proc. 48th IEEE Conf. Decision and Control, Shanghai, Dec. 2009, pp.
attitude stabilization,” in Proc. IFAC Symp. Nonlinear Control Systems, Tahoe 2522–2527.
City, CA, 1995, vol. 2, pp. 495–501. [40] H. K. Khalil, Nonlinear Systems. Englewood Cliffs, NJ: Prentice-Hall,
[15] P. Tsiotras and J. M. Longuski, “Spin-axis stabilization of symmetrical 2002.
spacecraft with two control torques,” Syst. Control Lett., vol. 23, no. 6, pp. [41] J. Cortés, “Discontinuous dynamical systems: A tutorial on solutions,
395–402, 1994. nonsmooth analysis, and stability,” IEEE Control Syst. Mag., vol. 28, no. 3,
[16] D. E. Koditschek, “The application of total energy as a Lyapunov func- pp. 36–73, 2008.
tion for mechanical control systems,” in Proc. AMS-IMS-SIAM Joint Sum-
mer Research Conf., Bowdin College, Maine, 1988, vol. 97, pp. 131–157. Dynam-

JUNE 2011 « IEEE CONTROL SYSTEMS MAGAZINE 51

Das könnte Ihnen auch gefallen