You are on page 1of 18

AIAA Guidance, Navigation, and Control Conference AIAA 2012-4845

13 - 16 August 2012, Minneapolis, Minnesota

Attitude control of satellites with flexible appendages:


a structured H∞ control design
∗ †
T. Loquen, H. de Plinval, C. Cumer and D. Alazard

In this paper, the modeling of flexible multi-body systems (typically a satellite), the
design of a reduced-order robust controller and the closed-loop system robustness analysis
are considered. First, we present a generic tool developed to build the model of a multi-
body dynamic system taking into account parametric uncertainties for robust control and
Downloaded by COLUMBIA UNIVERSITY on April 23, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4845

analysis purpose. Satellites are typically composed of a main body (hub of the satellite)
and cantilevered appendages (flexible or not). Based on the description of each element
(mechanical and geometrical properties, uncertain parameters ...), a linear model between
torques applied to the hub (control or perturbation inputs) and angular accelerations is
built using the Linear Fractional Representation framework. Second, based on this model,
a control design method is proposed: from a basic standard scheme and the latest H∞
solver, a low order H∞ controller can be derived. Roll-off requirements and worst-case
parametric configurations can also be taken into account in a multi-model design approach.
Third, robustness analysis of the design on the full order model is considered. A numerical
example illustrate the relevance and effectiveness of the proposed approach.

Nomenclature

MB subscript referring to the main body


A subscript referring to the appendage
G appendage center of gravity
P anchorage point of appendage on main body
B main body center of gravity
GFF main body reference frame (B, Bx , By , Bz )
SAR appendage reference frame (G, SAx , SAy , SAz )


F M B/A forces vector applied by the main body to the appendage


F M B/A,P torques vector applied by the main body to the appendage at point P

a→ linear acceleration vector of the main body at point P
P

−ω absolute angular velocity vector of the main body
vP linear velocity of the main body at point P
xP linear position of the point P
θ absolute angular position of the main body
S
DQ dynamic matrix of the subsystem S at point Q
η modal coordinates vector of flexible modes


F ext external forces applied to the main body


T ext,B torques vector applied to the main body at point B
τMi Mj geometric model between points Mi and Mj
T rotation matrix between SAR and GFF
∗ Research Scientists, ONERA, 2, Av. Edouard Belin, Toulouse- 31055 FRANCE.(Email: {thomas.loquen, henry.de plinval,

christelle.cumer}@onera.fr)
† Professor, University of Toulouse-ISAE, 10, Av. Edouard Belin, Toulouse- 31055 FRANCE. (Email: daniel.alazard@isae.fr)

1 of 18
Copyright © 2012 by ONERA - The French Aerospace Lab. Published by the American Institute of Aeronautics and Astronautics, Inc., with permission.
American Institute of Aeronautics and Astronautics
∆ matrix of uncertainties
∆wc matrix of worst case values of uncertainties
In identity matrix of order n
0n×m null n × m matrix
IG inertia matrix at point G
Ixx static gain of the system (angular model) around X axis
diag(A, B) matrix with elements A and B on diagonal, 0 otherwise
Subscript
i flexible mode number
Acronym
LFT Linear Fractional Transformation
dof degrees of freedom
AOCS Attitude Orbit Control System
Downloaded by COLUMBIA UNIVERSITY on April 23, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4845

I. Introduction
Spacecrafts are very complex mechanical multi-body systems including flexible and/or rotating ap-
pendages. Due to a lack of tools to deal with nonlinear spacecraft models, the design of the AOCS requires
a linear model taking into account all the rigid and flexible couplings between the hub (where the AOCS
acts) and the various appendages.
Note that the linear assumption is quite realistic for such systems since perturbations and motions are very
small (except for very agile observation satellites). This linear assumption is particularly valid in the context
of future missions for deep space exploration involving formation flying of several spacecrafts.
In this framework, engineers need a tool to develop quickly the dynamic model of the spacecraft including
the flexible modes, and eventually parametric uncertainties. In1 authors have proposed some tools developed
with MATLAB/SIMULINK to efficiently build the linear dynamic model of any open mechanical chain. The
method proposed in this paper is an extension of these tools in order to better study the attitude of the
satellite angular deflection around a pointing direction of the payload.
It must be emphasized that the multi-body modeling, presented in,1 first distinguishes the main body
dynamics from the flexible appendage dynamics and secondly connects them together. The resulting model
directly highlights the characteristic parameters of the main body and of the appendage. As a result, their
uncertainties can be explicitly introduced. Then a minimal LFR model of the satellite can be derived
and allows for robust stability to be analysed by classical methods (phase-gain margins, Bode plots and
µ-analysis2 ).
The second problem of interest considered in this paper addresses the control design problem. Indeed,
based on recent development of H∞ methods,34 and their implementation in MATLAB, structured and so
reduced-order H∞ controllers can be designed. Such new solvers (hinfstruct) present several advantages
with respect to classical full-order solvers already available
• the controller order is a design parameter. This is an advantage with respect to the classical H∞
approach where the controller order equals the augmented plant order, thus resulting in a complex
controller. As a result, all flexible modes can be taken into account in the design process, without the
need for a high order controller. Thus, in comparison with classical full order H∞ design, the model
and/or controller reduction steps are skipped,
• the design of a low order controller is a key point for implementation purpose
• beyond the controller order, the structured approach allows us to define one controller per axis thus
maintaining the axis decoupling property,

• multi-model H∞ design can also be performed.


Such a solver is applied on a standard control problem based on the acceleration sensitivity function,5 .6 This
basic scheme involves a reduced number of weighting parameters (by comparison with the standard 4-blocks
H∞ problem of 7 ). In addition, this standard problem is augmented to handle:

2 of 18

American Institute of Aeronautics and Astronautics


• roll-off specifications: the designed controller must be stable (strong stabilization) and must respect a
high frequency low-pass template,
• parametric worst-case configurations can be introduced into the design process (multi-model design).
The paper is organized as follows. In Section II, the multi-body modeling method of 1 is recalled and
completed to introduce satellite angular deflection modeling. In the third section, the control design problem
is addressed by considering incrementally standard problem more and more complex . Finally, a numerical
example is presented to emphasize the relevance of the proposed approach.

II. Multi body flexible mechanical system modeling


In this section we summarized the main results of 1 to derive the linear model of a rigid main body
connected to several appendages (flexible or not). This (inverse) dynamic model describes the relation
between the external forces and torques applied to the main body and the linear and angular accelerations
of the main body at the center of gravity B, as described in Fig. 1.
Downloaded by COLUMBIA UNIVERSITY on April 23, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4845

Figure 1. Linear Dynamic Model between forces and accelerations

II.A. Background
II.A.1. Definition and transport of the dynamic model of a rigid body
Let us consider a rigid body S, of mass m and inertia IG at its center of gravity G. The linear (direct)
dynamic model of the system S at point G is characterized by a block diagonal matrix:
" #
S mI3 03×3
DG = (1)
03×3 IG

such that (under linearity assumptions)


" → − →

# " #
F ext S aG

− = DG →
−̇ (2)
T ext,G ω

The dynamic model of S can be easily transported from its center of gravity G to a point P by the
relation:

DPS = τPT G DG
S
τP G (3)

where τP G is the geometrical model associated to points G and P:


" #
I3 (∗ GP )
τP G = (4)
03×3 I3

~ . If [x, y, z]T is the coordinates vector of


(∗ GP ) is the anti symmetric matrix associated to the vector GP
~ in the main body reference frame GFF, the expression of (∗ GP ) in this frame is:
GP
 
0 −z y
(∗ GP ) =  z 0 −x  (5)
 

−y x 0 GF F

3 of 18

American Institute of Aeronautics and Astronautics


II.A.2. Main body modeling
The direct dynamic DPM B of the main body (MB) at point P (the appendage anchorage point) can then be
written
" →− →
− →

# " # " # " #
F ext T mI3 0 aP aP

− = τBP τBP →
−̇ = DPM B →
−̇ (6)
T ext,P 0 IB ω ω

− →

where F ext and T ext,P denote the external forces and torques vector, respectively, applied to the satellite
at point P. Thus the transport of the direct dynamic model of a body MB from a point B to a point P reads:
" #
MB T MB mI3 m (∗ BP )
DP = τBP DB τBP = 2 (7)
−m (∗ BP ) IB − m (∗ BP )

If we consider now that an appendage A is cantilevered to the base MB at point P, the reaction force

− →

F M B/A and torque T M B/A,P at point P between the base and the appendage must be taken into account
Downloaded by COLUMBIA UNIVERSITY on April 23, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4845

in the dynamic model of the base.


" → − →
− →

# " #
F ext − F M B/A MB aP

− →
− = DP →
−̇ (8)
T ext,B − T M B/A,P ω

II.A.3. Appendage modeling



− →

It is assumed that the only force and torque applied to the appendage A are F M B/A and T M B/A,P . If
the appendage A is rigid, the principle is the same as for the main body modeling. But if the appendage
is flexible, its dynamic model is no more a static matrix. The appendage dynamic model MPA (s) can be
expressed under the following form:89 :
" →− →

# " #
F M B/A A aP

− = DP →
−̇ + LP η̈ (9)
T M B/A,P ω



" #
aP
η̈ + diag(2ξi ωi )ω̇ + diag(ωi2 )η = −LTP →
−̇ (10)
ω

This formulation depends on the appendage flexible modes ωi through three matrices: the stiffness matrix
K = diag(ωi2 ), the damping matrix D = diag(2ξi ωi ) and the modal participation factor matrix LP at point
P.

II.A.4. Connection of the main body and the flexible appendage


The whole satellite dynamics model gathers the equations related to the main body and to the flexible
appendage by use of the geometrical model τP B between points B and P and the rotation matrix T between
SAR (appendage reference frame) and GFF. Figure 2 shows that the direct dynamic model of the appendage
M B −1
at P interacts as a feedback on the inverse dynamic model [DB ] of the main body, where u and ωP are
respectively the control and the perturbation inputs. In case of many appendages other feedbacks are added
M B −1
on [DB ] , as shown by Figure 3 where the different subscripts k refer to the various appendages.

II.B. Angular deflection around a pointing direction of the payload


This subsection considers an antenna as a particular flexible appendage and describes the determination
of the absolute angular accelerations θ̈e and θ̈a of the pointing direction z~A around elevation axis x~A and
azimuth axis y~A , respectively, at the focal point F (see Figure 4).
To perform pointing error analysis on the antenna pointing direction, three new data are required to
described the antenna dynamic model:

4 of 18

American Institute of Aeronautics and Astronautics


Downloaded by COLUMBIA UNIVERSITY on April 23, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4845

Figure 2. Connection of dynamic models

Figure 3. Generalization of the model connection

• the coordinates of the focal point F in SAR

• the coordinates of the pointing direction in SAR


• the modal shape matrix φF (6 × N ) between the N modal coordinates and the 6 coordinates of the
displacements at point F in the reference appendage frame (SAR).

5 of 18

American Institute of Aeronautics and Astronautics


Then the model of the whole spacecraft involves two additional outputs corresponding to the absolute angular
accelerations at the point F (normal to the pointing direction, that is elevation and azimuth, and projected
in the SAR frame).
x
A

z
A
F

P
y
A

SA
R
B GFF
Downloaded by COLUMBIA UNIVERSITY on April 23, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4845

Figure 4. Sketch of the main body and the flexible antenna.

# "
~ F,SAR
δa
First, one can compute the 6 d.o.f. relative acceleration vector of the point F w.r.t. antenna
δ~ω̇F,SAR
reference frame SAR using the (6 × N ), N being the number of flexible modes of the antenna) modal shape
matrix φF between modal coordinates of the antenna η and the 6 displacements at point F . Considering
the state-space hybrid-cantilevered model MPA (s) of the antenna at point P :
       
η̇ 0k×k IN η 0k×6 ~aP
 =  +  
η̈ −Kk×k −D η̇ −LP ~˙
ω
     
F~B/A h i η ~aP
 = −LTP K −LTP D   + (DPA − LTP LP )   (11)
T~B/A,P η̇ ~˙
ω
Then: " #    
~ F,SAR
δa η ~aP
= φF η̈ = [−φF K − φF D]   − φF Lp   (12)
δ~ω̇F,SAR η̇ ~˙
ω
Then one has to take into account the acceleration of the SAR frame w.r.t. inertial frame at the point F :
" # " #  
~F
δa ~ F,SAR
δa ~aF
= +  (13)
δ~ω̇F δ~ω̇F,SAR ~˙
ω

with    
~aF ~aP
  = τF P   . (14)

ω ~˙
ω
Finally:    
" #
~F
δa η ~aP
= [−φF K − φF D]   + (τF P − φF Lp )   . (15)
δ~ω̇F η̇ ~˙
ω
Lastly:

θ̈e = < x~A , δ~ω̇F > (16)


θ̈a = < y~A , δ~ω̇F > , (17)

6 of 18

American Institute of Aeronautics and Astronautics


where < ~a, ~b > is the scalar product of vectors ~a and ~b.
All these equations are vectorial equations and thus frame independent. When implementing in the
modeling tool, all the vectors are expressed in the SAR. (x~A , y~A ) is an orthonormal basis of the plane
orthogonal to z~A (several solutions). In SAR, components [x~A , y~A ]SAR can be easily determined using the
null space of [z~A ]TSAR (that is: the row vector of pointing direction components in SAR):

[x~A , y~A ]SAR = ker([z~A ]TSAR ) . (18)

This way the dynamic model MpA (s) of the antenna is augmented with two additional outputs θ̈e and θ̈e
and is renamed MpA (s). Projecting in SAR this model reads:
2 3 2 32 3 2 32 3
η̇ 0k×k Ik η 0k×6 ~
aP
4 5=4 54 5+4 54 5
η̈ −Kk×k −Dk×k η̇ −LPk×6 ~˙
ω SAR

2 2 3 3
F~B/A
Downloaded by COLUMBIA UNIVERSITY on April 23, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4845

6 4 5 7 " # 2 3
6 7
LT
h i η
6 ~B/A,P
T 7
= − P (19)
SAR K D 4 5
kerT ([z~A ]T
6 7
SAR )φF (4 : 6, :) η̇
6 7
4 θ̈e 5
θ̈a
" #2 3
A
(DP − LT ~
aP
+ P LP ) 4 5 (20)
kerT ([z~A ]T
SAR )(I 3 − φF (4 : 6, :)Lp ) ~˙
ω SAR

The connection of this model M~pA (s) to the main body taking into account the (6 × 6) rotation matrix
T between SAR and GFF is depicted in Figure 5. Angular deflection can be determined by the double
integration of θ̈e and θ̈a . Note that these integrators are also non minimal and can be removed by computing
a minimal realization of the whole model between the 6 d.o.f force vector and the 6 + 2 position vector.
..
θe
..
θa
a
A P
MP (s) .
ω
SAR
T T
T

τT τ
PB PB

− MB −1 1 1
+ [ DB ] s s
F a
ext B
.
Text,B ω

Figure 5. Block Diagram of the Main body + antenna dynamic model with augmented outputs for pointing angular
deviation studies.

The next step consists in taking into account parametric uncertainties.

II.C. Global LFT model


Let us first recall that LFT (also called M (s) − ∆) framework is commonly used in order to exploit the well-
known µ-analysis tools for robustness analysis purpose. Such a framework uses multiplicative or additive
representation of uncertain parameters although in this study only multiplicative uncertainties are considered.

7 of 18

American Institute of Aeronautics and Astronautics


Figure 6. A M (s) − ∆ representation

The framework is based on isolating uncertainties in the matrix ∆ from the nominal system M (s), and then
connecting them through a feedback (see Figure 6).
To obtain the LFT model, one MATLAB LFR object per subsystem is created. This step is very simplified
Downloaded by COLUMBIA UNIVERSITY on April 23, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4845

by the fact that each parameter of the multi-body system have been introduced independently during the
modeling process and in a minimal way. Then, each uncertain parameter can be represented by an LFT
object in the structured model and the the LFT representation of the whole uncertain model can be easily
built by the interconnection of the various subsystems (see Figure 7 and also,1011 ).

Figure 7. Interconnection of LFR objects

III. Structured H∞ design


In this section, a control design method for multi-body flexible systems is presented. According to the
approach exposed in Section II, two linear dynamic models are assumed to be available:

• design model (DM): corresponding to the nominal inverse dynamic model (without uncertainty). This
model will be used to design the controller.
• validation model (VM): LFR model representing the actual system (with uncertainties) which will be
interconnected to the designed controller in order to perform the stability analysis.

8 of 18

American Institute of Aeronautics and Astronautics


One can note that, in this study, the DM includes all appendages flexible modes and avionics.

III.A. Standard problem definition


In this section, the standard problem definition is discussed. First the basic standard problem based on
the acceleration sensitivity function5 is presented. Then, an extended standard problem with controller
frequency shaping is proposed. Finally, a multi-model standard problem is defined taking into account a
worst-case configuration identified by a µ-analysis step.

III.A.1. A basic standard problem


The standard problem proposed in this paragraph allows to obtain a nominal design meeting the perfor-
mance requirements in terms of disturbance rejection. This standard problem involves very simple frequency
weighting functions directly defined by the specifications in terms of pulsation and damping ratio for each
degree of freedom. The proposed H∞ design weights the acceleration sensitivity function according to Figure
8. As before, double integrators between accelerations and positions are introduced considering the small
Downloaded by COLUMBIA UNIVERSITY on April 23, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4845

angle approximation.

Figure 8. A standard H∞ problem P(s) for disturbance rejection

The disturbance w acts on the acceleration vector, the controlled output z weights the acceleration with
the diagonal weighting matrix.
 s2 +2ξ ω s+ω2 
1 1
s2
1
0 ... ... 0
 .. .. 

 0 . 0 .



Wq =  .
.. s2
+2ξ ω
i i s+ω 2 .
..

(21)
0 i
0
 
s 2 
 .
.. . ..

0 0
 
 
2 2
s +2ξn ωn s+ωn
0 ... ... 0 s2

Wq (s) is the inverse of the desired acceleration sensitivity. For a degree of freedom i, the desired accel-
eration sensitivity function specifies that low frequency disturbances on the acceleration must be perfectly
rejected while high frequency disturbances (beyond the bandwidth ωi of this degree of freedom) are not re-
jected. The multi-variable weight Wq (s) is diagonal and depends only on the specifications ωi , ξi , i = 1 . . . n
and not on the system dynamic parameters.

III.B. The augmented standard problem


In the proposed approach, we use the multi channel feature of the MATLAB routine hinfstruct. In this
framework, the controller structure K (defined as the diagonal concatenation of controllers Kx , Ky and Kz )
becomes a decision variable of the optimization problem. This feature allows to add a channel for each
single axis controller to be optimized, as presented on Fig. 9. Such a scheme guarantees the stability of each
controller (strong stabilization) and allows to specify a roll-off behavior, directly on the frequency response
of the controller, through the templates WuX , WuY , WuZ . The first considered channel aims at minimizing
the H∞ norm of the transfer between the torque disturbance input w and the performance output z (see
Fig. 8). The other channels contain the product of the controller K on each coordinate (Kx, Ky and Kz)

9 of 18

American Institute of Aeronautics and Astronautics


1 1
with low pass filters WuX , WuY , W1uZ . The analytical formula for this control signal weighs is the following.
s
√ +1
Wu{X,Y,Z} = 2K{Ixx,Iyy,Izz} 100ωs cut (22)
ωcut + 1

where K{Ixx,Iyy,IzzZ} are the static gains needed for each coordinate to satisfy the correct disturbance
rejection. Note that these gains could correspond to those computed√ for a classical PD controller. The
pulsation ωcut is the roll-off filter cutting frequency. The weight 2 is introduced to add a 3dB margin. The
effect of theses templates on the frequency response of each controller is highlighted in Figures 15 to 17.
Note that this controller frequency-domain shaping is a quite original feature of this methodology, brought
by the structured H∞ optimization software. In particular, one must note that, on top of these achievements,
this methodology also ensures that no pole is added to cancel the system cantilevered zeros. Indeed, the
frequency shape ensures that there is no high peak in the controller frequency response, thus forbidding such
additional badly damped dynamics in the controller. This is a very desirable feature since such poles-zeros
cancellation is a well-known source of non-robustness in the closed-loop system.
Downloaded by COLUMBIA UNIVERSITY on April 23, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4845

Figure 9. Multi-channel H∞ problem

III.C. Design with worst-case uncertainty consideration


A multi-model scheme is now presented in order to obtain a more efficient controller if previous designs reveal
unsatisfactory considering parametric uncertainties. An iterative process is proposed which uses the capabilit
of the last hinfstruct release to include multi-model in the H∞ design and so worst cases parametric
configurations. More precisely, classical µ-analysis tools (computation of a µ-upper-and -lower bound) can
be used to identify a critical worst-case (i.e. once the µ-lower bound maximal value over frequency is greater
than one and the corresponding ∆-block value ∆wc is obtained), a new design procedure should be done by
integrating the worst case model to the H∞ optimization problem (see Fig. 10). As a result, the optimization
will provide an optimal controller from the point of view of the trade-off between the performance on the
nominal model and the performance on the worst-case model(s)

IV. Illustration
This section is dedicated to a numerical application of the proposed methodology. At his aim, we consider
a spacecraft composed of three main elements :

10 of 18

American Institute of Aeronautics and Astronautics


Downloaded by COLUMBIA UNIVERSITY on April 23, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4845

Figure 10. Multi-channel H∞ problem with worst case consideration

• a main body (MB)


• an antenna characterized by 22 flexible modes
• a solar array characterized by 10 flexible modes
The connection of these appendages to the main body, according to steps developed Section II leads to
64 × 64 model with 6 inputs and 6 outputs. In this examples, we only consider the 3 angular degrees of
freedom model: only the 3 last inputs and outputs will be considered.
The design of structured controller, i.e controller with fixed order, allows us to easily include dynamics
of avionics in the standard problem (see Figure 11). Actuators and sensors are modeled by second order low
pass filters with unit gain and bandwidths as proposed Table 2.

Reaction Wheels Gyrometers Star trackers


bandwidth (Hz) 100 200 10
Table 2. Avionics bandwidths

The standard problem in Figure 11 includes 88 states. Classical methods to design an H∞ controller
(Riccati-based or LMI-based H∞ optimization approach for example) lead to a full order controller. Beyond
the large computational time to obtain such a 88th order controller, a reduction is often needed, before
the synthesis if some flexible modes are supposed negligible or after the synthesis to obtain implementable
controller. Structured H∞ controller solvers allow a low order controller to be designed directly (in one
shot).

11 of 18

American Institute of Aeronautics and Astronautics


Figure 11. standard H∞ problem with avionics

IV.A. Nominal problem


Downloaded by COLUMBIA UNIVERSITY on April 23, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4845

In this section, we consider the nominal specifications. From the disturbance rejection requirements, the
weighting matrix (Eq. 21) can be easily tuned (see Table 3). The desired bandwidths of the closed-loop
system are very low w.r.t. the system first flexible mode frequencies. Then, this control problem is not
very challenging and a second order controller per axis is sufficient enough to meet requirement. Indeed, the
closed-loop performances are summarized Table 4. The corresponding Nichols plots are depicted Figure 12
and highlight good stability margins.

X axis Y axis Z axis


bandwidth (Hz) 0.1 0.1 0.1
damping ratio 0.7 0.7 0.7
Table 3. Nominal specifications

H∞ norm computational time (s)


Full order synthesis 1.0028 50.17
structured synthesis 1.01 28.3
Table 4. Synthesis performance

Regarding the time-domain simulation on the full order model, Figure 13 plots system responses to a
1deg initial conditions. The correct decoupling, provided by the block diagonal structure of the controller,
between all axes is illustrated by adding a torque perturbation of amplitude 5N m during 1s on X axis (at
time t = 80 s). The other axes are weakly perturbed by this impulsion.

IV.B. Challenging problem


Now the desired performances of the closed-loop system are arbitrarily increased to obtain faster time-
domain responses (see Table 5) and and to create dynamics couplings betwwen rigid modes and first flexible
modes. With such values, a second order controller is no more able to stabilize correctly the system (the

X axis Y axis Z axis


bandwidth (Hz) 1 1 1
damping ratio 0.7 0.7 0.7
Table 5. Challenging specifications

corresponding H∞ norm is too large). A solution consists in increasing the order of the controller. But in
such a case, we loose noise rejection benefits due to high frequency behavior of first or second order controller.

12 of 18

American Institute of Aeronautics and Astronautics


Downloaded by COLUMBIA UNIVERSITY on April 23, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4845

Figure 12. Nichols plots for X, Y, Z axes

Figure 13. Temporal responses to initial conditions and torque perturbation on X axis

Then roll-off specifications are required. The high frequency behavior of the structured controller can be
easily imposed by implementing the standard problem depicted Figure 9.
In this framework, the H∞ norm becomes 1.22 and the corresponding Nichols plot are proposed Figure
14.
On Figures 15, 16 and 17, frequency responses of controllers with (continuous black line) or without (con-
tinuous red line) roll-off constraints are plotted. The required frequency-domain behavior of the structured
controller can be verified axis by axis. Considering time-domain responses, the decoupling of each axes is
preserved. As expected, responses to initial conditions or torque perturbation are faster (see Figure 18).

13 of 18

American Institute of Aeronautics and Astronautics


Downloaded by COLUMBIA UNIVERSITY on April 23, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4845

Figure 14. Nichols plots for X, Y, Z axes

Figure 15. Frequency responses of designed controllers - X axis

IV.C. Robustness analysis


In this section, the robust stability analysis problem is addressed.
The modeling tool presented Section II.II.C allows us to build the LFR model of the spacecraft by con-
necting components with uncertain parameters. We emphasize that by construction, the obtained uncertain
model is minimal and no ∆ uncertainty block reduction can be done. The LFR system is an LFR-object
with 3 inputs, 3 outputs, 64 states and 168x168 ∆ uncertainty block. In this study case, the following data
have been considered subject to variations
• Main body
– inertias (9 parameters)
– center of gravity (3 parameters)
– mass (1 parameter)

14 of 18

American Institute of Aeronautics and Astronautics


Figure 16. Frequency responses of designed controllers - Y axis
Downloaded by COLUMBIA UNIVERSITY on April 23, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4845

Figure 17. Frequency responses of designed controllers - Z axis

• First flexible appendage (antenna)


– Main modal participations (13 parameters)
– pulsations (22 parameters)
– center of gravity (3 parameters)
– mass (1 parameter)
• Second flexible appendage (solar array)

– Main modal participations (4 parameters)


– pulsations (10 parameters)
– center of gravity (3 parameters)
– mass (1 parameter)

The corresponding uncertainties are detailed in Table 6.


The robust stability analysis is done by using the Skew-Mu Toolbox.2 In particular, if the obtained
µ-upper bound does not guarantee the robust stability of the uncertain system, a study of the µ-lower bound
can be done on a frequency gridding to find the worst-case parametric configuration.

15 of 18

American Institute of Aeronautics and Astronautics


Downloaded by COLUMBIA UNIVERSITY on April 23, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4845

Figure 18. Temporal responses to non null initial conditions and torque perturbation on X axis

Type MB Antenna Solar Array


center of gravity 12 18 16
(variation) (5%) (5%) (5%)
participation factor 26 8
(variation) (10%) (10%)
pulsation 44 20
(variation) (10%) (10%)
Inertias 9
(variation) (10%)
Mass 3 6 6
(variation) (10%) (10%) (10%)
Total 24 94 50
Table 6. Uncertainties repartition for robust performance analysis

Let us consider the 18th order controller designed section IV.B. The µ-lower bound analysis shows that
the system is not robustly stable with respect to uncertain parameter variations of Table 6: the maximum
µ-lower bound is greater than one on Figure 19.
This analysis provides a worst case value of the uncertain ∆-block. As proposed Figure 10, the new design
process should stabilize the systems in such a worst-case configuration, beyond the respect of constraints
of the previous design process. The synthesized controller leads to a H∞ norm of 1.24 for the closed-loop
system (very close to the one obtained section IV.B: 1.22). The robustness analysis of the interconnection of
this controller to the uncertain system is proposed Figure 20 and illustrates the effectiveness of the proposed
standard scheme: the maximal value of the µ lower bound decreased in a significant way.

V. Conclusion
A global methodology was proposed for multi-body dynamic systems -typically satellites- control and
analysis design. This method includes a generic dynamic modeling tool, including dynamic modeling for

16 of 18

American Institute of Aeronautics and Astronautics


Downloaded by COLUMBIA UNIVERSITY on April 23, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4845

Figure 19. µ-lower bound obtained with ωdes = [1 1 1]rad/s and without worst-case consideration

Figure 20. µ-lower bound obtained with ωdes = [1 1 1]rad/s and with worst-case consideration

control design and uncertainty representation for analysis purpose. The second step is a control design
method based on the latest version of the structured H∞ control design MATLAB function. Most impor-
tantly, this features low-order controller design, controller frequency shaping and worst case consideration.
Finally, more classical analysis tools are considered based on the uncertainty modeling earlier mentioned.
The whole method has been successfully applied to a generic satellite model as reported through numerical
simulation results.

Acknowledgements
The authors would like to thank Massimo Casasco from ESA and Chiara Toglia from Thales Alenia Space
Italy for their supports.

17 of 18

American Institute of Aeronautics and Astronautics


References
1 Alazard, D., Cumer, C., and Tantawi, K., “Linear dynamic modeling of spacecraft with various flexible appendages and

on-board angular momentums,” Proceeding of the 7th International ESA Conference on Guidance, Navigation and Control
Systems, Tralee (Ireland), June 2008.
2 Ferreres, G. and Biannic, J., Skew Mu Toolbox (SMT): a presentation, http://www.cert.fr/dcsd/idco/perso/Ferreres/index.html,

2003.
3 Apkarian, P. and Noll, D., “Nonsmooth H
∞ synthesis,” IEEE Transactions on Automatic Control,, Vol. 51, No. 1, jan.
2006, pp. 71 – 86.
4 Gahinet, P. and Apkarian, P., “Structured H
∞ Synthesis in Matlab,” IFAC World Congress, Università Cattolica del
Sacro Cuore, Milano, Italy, 2011.
5 Alazard, D., Fezans, N., and Imbert, N., “Mixed H /H
2 ∞ control design for mechanical systems: analytical and numerical
developments.” AIAA Guidance Navigation and Control conference, Honnolulu (Hawaı̈¿ 12 ), August 2008.
6 Guy, N., Alazard, D., Cumer, C., and Charbonnel, C., “Reduced Order H-Infinity Controller Synthesis for Flexible

Structures Control,” 7th IFAC Symposium on Robust Control Design, Aalborg, Denmark,, June 2012.
7 Charbonnel, C., “H
∞ and LMI attitude control design: towards performances and robustness enhancement,” Acta
Astronautica, Vol. 54, No. 5, 2004, pp. 307 – 314.
8 Cumer, C. and Chretien, J., “Minimal LFT form of a spacecraft built up from two bodies.” Proceedings of the 2001

AIAA Guidance, Navigation, and Control Conference and Exhibit, AIAA, 2001.
Downloaded by COLUMBIA UNIVERSITY on April 23, 2013 | http://arc.aiaa.org | DOI: 10.2514/6.2012-4845

9 Fezans, N., Alazard, D., Imbert, N., and Carpentier, B., “H


∞ control design for multivariable mechanical systems -
Application to RLV reentry.” AIAA Guidance Navigation and Control conference, Hilton Head, August 2007.
10 Biannic, J. and Doll, C., Introduction to a Simulink-based interface for LFRT Toolbox ,
http://www.cert.fr/dcsd/idco/perso/Magni/m simulink/index.html, July 2009.
11 Magni, J., User Manual of the Linear Fractional Representation Toolbox Version 2.0 ,
http://www.cert.fr/dcsd/idco/perso/Magni/index.html, July 2009.

18 of 18

American Institute of Aeronautics and Astronautics

You might also like