Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

AbstractFor an air-breathing hypersonic vehicle model

subject to input constraints, this paper is concerned with the


application of linear control theory to design its flight control
system. In the framework of integral quadratic constraints
(IQCs), the standard linear parameter-varying (LPV) control
method is generalized to deal with the parametric, dynamic
uncertainties and saturation nonlinearities associated with the
model. Based on open-loop analysis, an inner/outer loop
structure is proposed to design the flight control system that is
gain-scheduled over dynamic pressure and Mach number. The
inner-loop consists of robust LPV controller and LPV
anti-windup compensator to enhance attitude stability and
overcome actuator saturation. The outer-loop is designed as a
gain-scheduled proportional-integral (PI) controller to achieve
precise trajectory tracking performance. The effectiveness of
the control strategy is demonstrated on the full nonlinear model.
I. INTRODUCTION
ONTROL of air-breathing hypersonic vehicles is a
significant challenge due to the strong coupling effects
between aerodynamics, propulsion, and structure. The
vehicles are statically unstable in pitch mode, and exhibit
non-minimum phase behavior. The dynamic characteristics
of vehicles vary significantly due to the wide range of flight
conditions, fuel consumption, and thermal effects on the
structure. In addition, significant uncertainties affect the
models. Reference [1] presents a thorough survey of the
difficulties in the modeling and control of hypersonic
vehicles.
As the linear control designs surveyed in [1] and recent
literatures [2]-[4] apply only to a certain flight condition, the
gains of the controllers should be scheduled to be compatible
with changing operating conditions. As a good starting point
for analysis and design of more general gain scheduling
problems, LPV systems received a growing interest in the
control community in recent years. The synthesis problem is
often shown to be equivalent to the solution of some
parameter-dependent linear matrix inequalities. The standard
control synthesis method is summarized in [5] in detail.
The motivation for this research is twofold. Firstly, to deal
with the modeling uncertainties and input saturation

Manuscript received October 9, 2009. (Write the date on which you
submitted your paper for review.) This work was supported by the National
Natural Science Foundation of China under Grant 60874084.
Dong Ming. Ge is with the Harbin Institute of Technology, Harbin China
(phone: 0086-0451-86402204-8666; e-mail: gedm1982@ 163.com).
Xian Lin. Huang is with the Harbin Institute of Technology, Harbin China
(phone: 0086-0451-86402204-8666; e-mail: xlinhuang@ hit.edu.cn).
nonlinearities associated with the vehicle model, we extend
the standard LPV method in [5]. In the framework of integral
quadratic constraints, control synthesis algorithm is presented
in terms of scaled linear matrix inequalities. Secondly, based
on LPV modeling, the developed method is used to design
robust gain-scheduled flight control system. According to the
property of two time scales between attitude dynamics and
trajectory dynamics
[6]
, a two-loop control architecture is
adopted. The inner-loop control consists of robust LPV
controller and LPV anti-windup compensator to enhance
attitude stability and overcome actuator saturation. The
outer-loop control is designed as gain-scheduled
proportional-integral compensator to guarantee precise
trajectory tracking performance. Nonlinear simulation results
are presented to demonstrate the effectiveness of the
theoretical results and application.
The notations are defined as follows. Let
n n
A denotes
n-dimensional diagonal matrix. A block-diagonal structure
with sub-blocks X
1
, X
2
,, X
p
on its main diagonal is denoted
as diag(X
1
, X
2
,, X
p
). sy(AS) is equal to AS+S
T
A
T
. L
n
2e
denotes
n-dimensional functional space whose members only need to
be square integrable on finite intervals.
r
w
r
z
p
w
p
z
u y
( ) P
A
( ) K

Fig. 1. Standard LFT control structure.
II. ROBUST LPV CONTROL METHOD
A. System Description Based on IQCs
The concerned problem is formulated as the linear
fractional transformation (LFT) structure shown in Fig. 1.
The block P() denotes the generalized LPV plant of the form
(1), and typically includes the controlled LPV plant together
with weighing functions specified by the user.
( ) ( ) ( )
( ) ( ) ( )
( ) ( ) 0
p p w u p
z zw zu
y yw
x A B B x
z C D D w
y C D u



(
( (
(
( (
=
(
( (
(
( (

(1)
Here, [ ]
T T T
r p
w w w = and [ ]
T T T
r p
z z z = .
n
p
x e9 are the states.
Robust Linear Parameter-varying Control of Air-breathing
Hypersonic Vehicle Considering Input Constraints
Dong Ming. Ge, Xian Lin. Huang
C
2010 8th IEEE International Conference on
Control and Automation
Xiamen, China, June 9-11, 2010
WeD2.3
978-1-4244-5196-8/10/$26.00 2010 IEEE 279



The input/output channels associated with the robustness and
performance criterion are
r
w ,
r
n
r
z e9 and
p
w ,
p
n
p
z e9 ,
respectively.
u
n
u e9 are the control inputs, and
y
n
y e9 are
the measurement outputs. The parameter vector are
assumed to be measurable online and contained in set (2)

{ }
1
: ( ) ( , ) : , , 1, ,
s
i i
S t C i s

v
+
= e 9 9 e s = ! ( (2)
where (c9
s
is a compact set.
is a causal operator from L
r
2e
|0,) to L
r
2e
|0,) with its
inputs and outputs satisfying the following time-domain
integral quadratic constraint
t
0
( ) ( )
0 0.
( ) ( )
T
T
r r
r r
w t w t
Q S
dt t
z t z t
S R
( ( (
> >
( ( (

)
(3)
Let Q, S, R be scaling matrices such that Q<0, R>0. We
assume that is block-diagonal: =diag(
1
,,
r
), where
i

denotes a troublemaking component. Note that the bounded
induced L
2
norm performance
2 2
p p
z w s can also be
described by IQC. The IQC characterizations for the typical
cases considered here are listed in Table. 1. The
sector-bounded nonlinearity in Table. 1 satisfies =I-,
where denotes standard saturation operator. For application,
all of the individual IQC are collected in block-diagonal
matrices Q=diag(Q
1
,, Q
r
, Q
p
), R=diag(R
1
,, R
r
, R
p
), and
S=diag(S
1
,, S
r
, S
p
) to characterize the associated
composition of and induced L
2
norm performance.
TABLE I
IQC CHARACTERIZATION FOR SPECIFIED
i
A AND PERFORMANCE
[0, ] K
u
ue ( )
i
i n
t I o
1
diag( , , )
i
n
o o ! ( ) 1
i
s

A < Type
2
i i n n
i
Q V

= eA
i
i n
R I c =
i
S VK
u
=
i i n n
i
Q

e9
i i
R Q =
0
i
S =
i
i n
Q qI =
i i
R Q =
i i n n
i
Q

eA
0
T
i i
S S + = 0
i
S =
i i
R Q =
Scalings
2 2
p p
z w s
p p n
Q I =
0
p
S =
(1/ )
p
p n
R I =

Consider the LPV system (1), we assume that a full-order
dynamic LPV controller K() is of the form
( , ) ( )
,
( ) ( )
k k k k
k k
A B x x
C D u y


( ( (
=
( ( (


(4)
where x
k
e9
n
is the controller state. Then, the final
closed-loop system admits the realization
( , ) ( )
,
( ) ( )
c c c c
c c
A B x x
C D z w


( ( (
=
( ( (


(5)
where x
c
=|x
T
p
x
T
k
|
T
. The gain-scheduled output feedback
control problem consists of finding the dynamic LPV
controller (4) such that the closed-loop system (5) is robustly
stable with respect to the perturbation block while
minimizing induced L
2
norm bound for all (t)eS

. For
simplicity, the dependence of data and variables on and
has been dropped somewhere.
B. System Analysis Condition
Theorem 1 (Analysis condition). Given >0, if there exist
parameter-dependent symmetric matrix P() and scaling
matrices Q(), R(), S() such that P()>0 and
0
T T T
c c c c c
T T T T T
c c c c c
c c
A P PA P PB C S C R
B P S C Q S D D S D R
RC RD R
(
+ + +
(
+ + + <
(
(

(6)
holds for all (t)eS

, then the closed-loop system (5) is robust


stable against the uncertainty block satisfying the IQC (3),
and has the induced L
2
norm performance bound .
Proof. Given the system (5), we assume a parameter
-dependent quadratic function V(x
c
)=x
T
c
P()x
c
such that
0.
T T T T T
V w Qw w S z z Sw z Rz + + + + <

(7)
The inequality (7) is rewritten as
0
( ( ) ) 0.
t
T T T T T
d
V w Qw w S z z Sw z Rz d
dt
t + + + + <
)
(8)
Note that the second term is always nonnegative. According
to standard arguments from Lyapunov theory, the system is
stable. Here, the function V decreases to zero, but not
necessarily monotonically. Thus, V is not a Lynpunov
function in the conventional sense. Condition (7) is
equivalent to (6) by Schur complement.
C. Control Synthesis
Based on the analysis condition, in this subsection we aim
to present a constructive procedure to design the controller
(4). With the controller system matrices notated as
,
k k
k
k k
A B
C D
(
O =
(

(9)
Theorem 1 can be reformulated as
0
T T T
k k
Z E F F E + O + O < (10)
by the fact that the O
cl
of the closed-loop system (5) has affine
dependence on O
k
. Partitioning P and P
1
as
1
,
T T
Y N X M
P P
N M

( (
= =
( (
- -

(11)
with X, Y, M, Ne9
nn
. Projection Lemma
[7]
leads to the
following solvability conditions for controller O
k
.
Theorem 2 (Solvability conditions). Consider the LPV
plant of the form (1) with scheduled parameter vector (t)
defined as (2), and let N
X
and N
Y
denote orthonormal bases of
the null spaces of |B
T
u
, D
T
zu
R, D
T
zu
S| and [C
y
, D
yw
] respectively.
If there exist parameter-dependent symmetric matrices X(),
Y()e9
nn
satisfying the following inequalities for all
(t)eS

, then there exists an nth-order LPV controller O


k
to
robustly stabilize the closed-loop system and has the induced
L
2
norm bounded by .
( )
0
( )
T T
p z w z
T
X z zw X
T T T T
w z zw zw
sy A X X XC R B XC S
N RC X R RD N
B S C X D R Q sy D S
(
+
(
<
(
(
+ +


(12)
WeD2.3
280



( )
( ) 0
T T
p w z z
T T T T T
Y w z zw zw Y
z zw
sy YA Y YB C S C R
N B Y S C Q sy D S D R N
RC RD R
(
+ +
(
+ + <
(
(


(13)
0, ,
r p
Y
X X Y
n n
N
X I
N N N
I
I Y +
(
(
> = =
(
(
(


(14)
After solving X(), Y() matrix functions, the LPV
controller (4) can be constructed explicitly by the following
Theorem, which is an extension of the results in [7].
Theorem 3 (Controller construction). Given X, Y and
scaling matrices Q, S, R, a LPV controller O
k
of order n can
be constructed along the following steps.
1) Compute invertible matrices M, Ne9
nn
via singular
value decomposition such that MN
T
=I-XY.
2) Compute D
k
such that
0
T T T
c c c
c
c
Q S D D S D R
RD R
(
+ +
A = >
(


(15)
c zw zu k yw
D D D D D = + (16)
3) Compute
k
B

and
k
C

solutions to the following linear


matrix equations.
0 0
0
yw y
T
T T T T T T T k
yw c c c w z zu k y
c z zu k y
D C
B
D Q S D D S D R B Y S C S D D C
RD R RC RD D C
( (
( ( (
+ + = + +
( ( (
-

( (
+


(17)
0
( )
T T T
zu zu u
T T T T T T k
zu c c c w u k yw z
zu c z
D S D R B
C
S D Q S D D S D R B B D D XC S
RD RD R RC X
( (
( ( (
+ + = + + ( ( (
-
( (

( (


(18)
4) Compute A
k
, B
k
, C
k
by the following equations
k k u k
NB B YB D =

(19)
T
k k k y
C M C D C X =

(20)
1
( )
T
T T T T
w u k yw z k zu
z zu k
B B D D XC S C D S
L
RC X RD C
(
+ + +
=
(
+

(21)
2
( )
T
T T T T T
w k yw z y k zu
z zu k y
YB B D C S C D D S
L
RC RD D C
(
+ + +
=
(
+
(

(22)
1
2 1
( )
( )
T
k u k k y p u k y
T T T
p u k y c
NA M YB C B C X Y A B D C X
A B D C YX NM L L

= + +
+ + + A


(23)
Due to the fact that the matrix variables X, Y and scaling
matrices Q, S, R enter inequalities (12-13) in nonlinear
fashion, we have to resort to the following D/K-like iterative
algorithm based on LMI.
1) Initialize scaling matrices Q, S, R.
2) With fixed Q, S, R, perform control synthesis of O
k

according to Theorem 2 and Theorem 3.
3) Apply Theorem 1 to the closed-loop system (5) to solve
scaling matrices Q, S, R minimizing .
4) Iterate over step 2) to step 3) until can not be decreased
significantly.
In fact, the Theorem 1-3 involve the infinite-dimensional
LMI optimization problem. This difficulty can be overcome
by parameterizing the matrix variables and gridding the
parameter space
[5]
. It must be stressed out that the controller
depends as well on the variable , which is maybe
prohibitive in practice. In our application, a single quadratic
Lyapunov function V(x
c
)=x
T
c
Px
c
is chosen in the LPV control
synthesis. This allows us to obtain gain-scheduled controller
not depending on and reduce computation time.
III. APPLICATION TO HYPERSONIC VEHICLE CONTROL
A. Air-breathing Hypersonic Vehicle Model
The longitudinal dynamics for the widely used
air-breathing hypersonic vehicle model
[8]
considered in this
paper, is described by the following nonlinear equations
2
2
2
( cos )
sin
( sin ) ( )
cos
sin
,
yy
E
T D
V
m r
T L V r
mV Vr
M
q q h V
I
r h R
o
u
o
u u
0 u
o 0 u

=
+
=
= = =
= = +

(24)
where V, u, q, 0, h, o, , R
E
are speed, flight path angle, pitch
rate, pitch angle, altitude, angle of attack, gravity constant
and earth radius, respectively. Control inputs are fuel
equivalence ratio |e[0.1 1.2] and control surface deflection
o
e
e[-20 20] deg. Control actuator dynamics are represented
by first order filters with bandwidth 100 rad/sec for | and 30
rad/sec for o
e
. Following [9], a curve-fitted model (CFM) has
been derived for control design. The approximations of the
forces and moments are given as follows
,
( , , ), ( , , )
[ ( , ) ( , )]
( , , )
L e D e
T T
T M e
L qSC Ma D qSC Ma
T qS C Ma C Ma
M z T qScC Ma
|
o o o o
o | o
o o
~ ~
~ +
~ +
(25)
In contrast to [9], the force and moment coefficients of the
CFM depend on the Mach number Ma. Note also that the
effect of the dynamic pressure q on thrust is included.
B. LPV Model
To conduct gain-scheduling control system design, the
nonlinear model should be transformed into a LPV model.
Through Jacobian linearization along the equilibrium point
parameterized by (t), the LPV model is developed as
[ ( )] [ ( )] x A t x B t u = + (26)
with state variables x=[V, u, q, 0]
T
and control variables u=[|,
o
e
]
T
. The outputs to be controlled are y=[V, u]
T
. For simplicity,
the variables represent perturbations from their equilibrium
states when linearization is considered. (t)=[ q , Ma]
T
e
WeD2.3
281



[24000 96000] pa [7 12] are used as scheduling parameters,
which capture the variability of the vehicle dynamic
characteristics with flight conditions.
Fig. 2 shows the distribution of the open-loop poles at the
vertices of parameter varying intervals. As expected, the
vehicle behavior depends significantly on the flight
conditions. The vehicle would be subject to attitude
divergence due to an unstable short-period mode. Note also
that the amplitudes of the phugoid modes are far below those
of the short-period modes. This is the so-called path-attitude
decoupling phenomenon analyzed in [6]. It is exactly this
two time-scale property justifying the dual loop control
architecture adopted below.
-2 -1.5 -1 -0.5 0 0.5 1 1.5 2
-0.015
-0.01
-0.005
0
0.005
0.01
0.015
Real Axis
I
m
a
g
in
a
r
y

A
x
is
Phugoid Mode
Short-period Mode

Fig. 2. Open-loop system poles for the extremal values of q and Ma.
C. Inner-loop Control with Anti-windup Compensation
The inner-loop control acts to stabilize the attitude
dynamic modes, and track pitch angle command issued by the
outer-loop controller. Because air-breathing vehicles are
vulnerable to the angle of attack, small changes in angle of
attack can translate into significant pitching moments. This
can easily lead to overlarge control deflections, and even
actuator saturation. Applying the developed LPV method in
section 2, the inner-loop control designed as robust LPV
controller augmented with LPV anti-windup compensator.
The short-period LPV model is selected from (26) as
( ) ( ) ( )
in in e
q A q B
u u
o
0 0
(
(
(
(
= + +
(
(
(
(

(27)
Firstly, a robust LPV controller is designed ignoring the
saturation operator . The parametric uncertainties originate
from the derivatives of lift L and pitching moment M with
respect to angle of attack o and control surface deflection o
e
,
with uncertainty level of 20%. To avoid excitation of elastic
modes, a multiplicative input uncertainty A
d
(s) with the cover
function W
d
(s)=2.5[(s+2)/(s
2
+9.76s+381.4)] taken from [10]
is placed at the control input. With the parametric
uncertainties extracted from the system in LFT way and
rescaled to [-1 1], we end up with the control interconnection
shown in Fig.

3.
Plant
l
I o
m
I o
d
W
p
z
p
w
2 r
w
1 r
w
1 r
z
2 r
z
q
0
LPV
Controller
W
0
e
W
d
A
Act
e
o


Fig. 3. Control interconnection for robust LPV controller.
The control problem is formulated as a model following
problem. Corresponding to varying operating conditions, we
select the following parameter-varying reference model
2
2 2
( , )
2 ( , ) ( , )
w q Ma
W
s w q Ma s w q Ma
0

=
+ +
(28)
A damping factor of =0.8 appears adequate for most flight
conditions. The value of natural frequency ranges linearly
from 4 rad/sec at lowest dynamic pressure and highest Mach
number to 6 rad/s at highest dynamic pressure and lowest
Mach number. The weighting function is chosen as a
low-pass filter W
e
=20[(0.1s+1)/(10s+1)]. With the parametric
uncertainties and dynamic uncertainty combined as a
block-diagonal uncertainty structure A=diag(o
l
I, o
m
I, A
d
), the
control diagram in Fig. 3 is equivalently reformulated as the
LFT structure in Fig. 1 for design. Solving the LMIs is done
by parameterizing matrix variables and scaling matrices as
affine parameter-dependent structure and gridding the
two-dimensional parameter space
[5]
. Applying the developed
synthesis method, we achieve =1.6.
An anti-windup controller proceeds to be designed in a
similar way. For exponentially unstable plant, only local
stability can be obtained with bounded inputs. The control
interconnection for anti-windup controller is shown in Fig. 4.
We combine the sector bounded nonlinearity =I- with
modeling uncertainties as a block-diagonal uncertainty
structure A=diag(, o
l
I, o
m
I, A
d
). The weighing function W
e
is
selected as constant weight W
e
=0.01. K
u
=0.7 is adopted to
allocate the partial design freedom for coping with robustness
and performance. Then, the anti-windup control diagram in
Fig. 4 is equivalently reformulated as the LFT structure in Fig.
1 for design. As a result, we achieve =38.6. The guaranteed
stability region corresponds to the operating range of control
defection |o
e
|s(1/(1- K
u
))20 deg.
Plant
+
LPV Anti-windup
Compensator
e
W
p
z
p
w 1 r
w
1 r
z
q
0
LPV
Controller
d
W
2 r
w
2 r
z
d
A
l
I o
m
I o
1 r
z
Act
e
o


Fig. 4. Control interconnection for LPV anti-windup compensator.
WeD2.3
282



D. Outer-loop Control
The outer-loop controller commands fuel equivalence ratio
and pitch angle to track velocity and flight path angle. The
engine and inner-loop dynamics are ignored in design
because the outer-loops bandwidth is much lower than the
engine and inner-loop dynamics. Replacing the control
deflection o
e
in (26) with o
e,trim
=-(A(3,4)o+B(3,1)|)/B(3,2)
which provides a zero pitching moment, the outer-loop
dynamics model is developed as
( ) ( )
out out
V V
A B
|

u 0 u
( ( (
= +
( ( (

(29)
Here, the outputs to be controlled are given by [V, u]
T
. To
guarantee precise trajectory tracking, integrators are
considered to charge up the input values required to trim
the vehicle. Hence, the out-loop control is designed as
gain-scheduled proportional-integral controller
( )
,
( )
ref
P I
ref
V V dt
V
K K
dt
|
0 u
u u
(

( (
(
=
( (
(


)
)
(30)
where V
ref
and u
ref
are reference commands. Define the
closed-loop state vector as
[ ( ) , ( ) , , ]
T
PI ref ref
x V V dt dt V u u u =
) )
(31)
then, the closed-loop system admits the realization
2 2 2
2
ref
PI PI
ref out I out out P
V O I I
x x
B K A B K O u
( ( (
= +
( ( (


(32)
where O
2
denotes the 22 zero matrix, and I
2
denotes the 22
identity matrix. Now select K
I
and K
P
to achieve the desired
closed-loop dynamics
2 2
11 21 12 22
diag( , ) diag( , )
ideal
O I
A
a a a a
(
=
(


(33)
where, for time-invariant closed-loop dynamics, a
i1
>0, a
i2
>0,
i=1, 2, are the coefficients of the desired closed-loop
characteristic polynomial of each channel given by

2
+a
i2
+a
i1
. As a result, we obtain
1
12 22
1
11 21
( diag( , ))
diag( , )
P out out
I out
K B A a a
K B a a

=
=
(34)
IV. SIMULATION EVALUATION
To validate the controllers designed in the above section,
this section presents nonlinear simulations performed with
the widely recognized nonlinear model
[5]
and standard
atmosphere model in Simulink. All simulations used fixed
step Runge-Kutta integration with a step size of 0.005 sec. To
take into account the aerodynamic/propulsive data error,
force and moment coefficients have been selected with a
20% variation of the nominal values. The nonlinear
simulation diagram is schemed in Fig. 5.
The selected maneuvering reference trajectory is referred
to the one defined in [9], which is considered to be a plausible
Flight Vehicle Model Actuator
Inner-loop
Controller
Outer-loop
Controller
Reference
Trajectory

cmd
0
0
q
e
o
|
[ , ]
T
q Ma
ref
ref
V
u
(
(

V
u
(
(


Fig. 5. Nonlinear simulation diagram.
operating trajectory for the vehicle, and the wide range of
operating conditions provide a moderately aggressive
tracking challenge to the gain-scheduling control. For limited
space, specific definition is referred to [9]. Next, simulation
results for two illustrative case studies are presented,
corresponding to tracking problem and anti-windup problem,
respectively.
0 50 100 150 200 250 300 350 400 450
2300
2400
2500
2600
2700
2800
2900
3000
3100
(a) Time (s)
V
e
lo
c
ity
(m
s
-1
)


V
Vref
0 50 100 150 200 250 300 350 400 450
-0.1
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
(b) Time (s)
F
lig
h
t P
a
th
A
n
g
le
(d
e
g
)


u
uref

0 50 100 150 200 250 300 350 400 450
1
1.5
2
2.5
3
3.5
4
4.5
5
(c) Time (s)
A
n
g
le
o
f A
tta
c
k
(d
e
g
)
0 50 100 150 200 250 300 350 400 450
1.5
2
2.5
3
3.5
4
4.5
5
(d) Time (s)
P
itc
h
A
n
g
le
(d
e
g
)


0
0cmd

0 50 100 150 200 250 300 350 400 450
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
(e) Time (s)
E
q
u
iv
a
le
n
c
e
R
a
tio
0 50 100 150 200 250 300 350 400 450
8
9
10
11
12
13
14
15
16
17
(f) Time (s)
E
le
v
a
to
r (d
e
g
)

Fig. 6. Nonlinear response for tracking the reference trajectory.
In the first simulation study, the vehicle is controlled to
track the reference trajectory mentioned above. Fig. 6 shows
the vehicle response for the test trajectories under the
designed controllers. Specifically, the tracking performance
for the velocity and flight path angle is excellent (see Fig. 6a
and Fig. 6b). The behavior of the angle of attack is given in
Fig. 6c, and the inner-loop controller behaves good tracking
performance for the pitch angle commanded by the
outer-loop controller (see Fig. 6d). Finally, Fig. 6e and Fig. 6f
show the control curves of fuel equivalence ratio and control
surface deflection, which range within their bounds. The
results of the simulation confirm that the gain-scheduling
control provides stable tracking of the reference trajectories
during the entire maneuver and exhibits good robustness to
modeling errors. Note also that the nearly perfect tracking
performance can be attributed to the employed inner/outer
control strategy which separately addresses fast tracking and
zero steady-state error concerns.
The second case study aims to illustrate the effectiveness
WeD2.3
283



of the anti-windup scheme. Although high maneuvering
flight in hypersonic speed is undesirable and low frequency
reference tracking is usually used to keep the inputs in their
linear operating ranges. However, unexpected atmospheric
disturbances can still result in transient changes in angle of
attack, causing actuator saturation and instability in attitude.
This consideration is exactly the purpose of the introduction
of the anti-windup compensation in control design. Similarly,
the vehicle is controlled to tracking the reference trajectory.
However, the vehicle is subjected to two impulse
perturbations of 4 deg with duration of 10 sec in angle of
attack at 100 sec and 300 sec, which are representative of
abrupt atmospheric disturbance. Fig. 7 shows the stable
response of the vehicle under actuator saturation. The
zoomed saturation curve of elevator is shown in Fig. 7g and
Fig. 7h. During the saturation period, the anti-windup
compensator seeks to modify the control law to drive the
elevator back to linear operating range. After the control
signal moves out of the saturation zone, the robust LPV
controller resumes its full tracking capability. However,
when the anti-windup compensator is unused, the elevator is
saturated severely and the vehicle goes unstable. For limited
space, the response is not given. Hence, anti-windup control
could actually augment the domain of attraction of the control
system and enhance its capability of disturbance rejection.
0 50 100 150 200 250 300 350 400 450
2300
2400
2500
2600
2700
2800
2900
3000
3100
(a) Time (s)
V
e
lo
c
ity
(m
s
-1
)


V
Vref
0 50 100 150 200 250 300 350 400 450
-0.2
0
0.2
0.4
0.6
0.8
1
1.2
(b) Time (s)
F
lig
h
t P
a
th
A
n
g
le
(d
e
g
)


u
uref

0 50 100 150 200 250 300 350 400 450
-5
-4
-3
-2
-1
0
1
2
3
4
5
(c) Time (s)
A
n
g
le
o
f A
tta
c
k
(d
e
g
)
0 50 100 150 200 250 300 350 400 450
-5
-4
-3
-2
-1
0
1
2
3
4
5
(d) Time (s)
P
itc
h
A
n
g
le
(d
e
g
)


0
0cmd

0 50 100 150 200 250 300 350 400 450
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
1.1
1.2
(e) Time (s)
E
q
u
iv
a
le
n
c
e
R
a
tio
0 50 100 150 200 250 300 350 400 450
-5
0
5
10
15
20
25
(f) Time (s)
E
le
v
a
to
r (d
e
g
)

98 100 102 104 106 108 110 112 114 116
-5
0
5
10
15
20
25
(g) Time (s)
E
le
v
a
to
r (d
e
g
)
298 300 302 304 306 308 310 312 314 316
-5
0
5
10
15
20
25
(h) Time (s)
E
le
v
a
to
r (d
e
g
)

Fig. 7. Nonlinear response with LPV anti-windup compensator under
disturbance.
V. CONCLUSION
The dynamic response characteristics of longitudinally
maneuvering air-breathing hypersonic vehicles vary
substantially with flight conditions. These changes require
scheduling of the flight control system to be compatible with
changing operating conditions. For dealing with the modeling
uncertainties and saturation nonlinearities associated with the
vehicle model, we extended the standard LPV control
algorithm. Based on LPV modeling of the nonlinear vehicle
model, the extended LPV method is applied to design a
gain-scheduling flight control system. The controller design
combines a pair of robust LPV controller and LPV
anti-windup compensator and a gain-scheduled proportional
-integral controller as inner/outer loop structure, thereby
achieving attitude stability and precise trajectory tracking
performance. The introduced LPV anti-windup compensator
can successfully overcome the actuators momentary
saturation caused by abrupt atmospheric disturbance. The
nonlinear simulation results appear to be sufficiently
promising to warrant further study of the developed LPV
control algorithm and application to flight control system
design.
REFERENCES
[1] B. Fidan, M. Mirmirani, and P. A. Ioannou, Flight dynamics and
control of air-breathing hypersonic vehicles: review and new
directions, AIAA Paper No. 03-7081, Proceedings: AIAA
International Space Planes and Hypersonic Systems and Technologies
Conf., Norfolk, Virginia, December, 2003.
[2] K. P. Groves, D. O. Sigthorsson, and A. Serrani, el al, Reference
command tracking for a linearized model of an air-breathing hypersonic
vehicle, AIAA Paper No. 05-6144, Proceedings: AIAA Guidance,
Navigation and Control Conf., San Francisco, California, August,
2005.
[3] M. Kuipers, M. Mirmirani, P. Ioannou, and Y. Huo, Adaptive control
of an aeroelastic airbreathing hypersonic cruise vehicle, AIAA Paper
No. 07-6326, Proceedings: AIAA Guidance, Navigation and Control
Conf., South Carolina, August, 2007.
[4] D. O. Sigthorsson, P. Jankovsky, A. Serrani, el al, Robust linear output
feedback control of an airbreathing hypersonic vehicle, Journal of
Guidance, Control, and Dynamics, vol. 31, pp. 10521066,
July-August 2008.
[5] P. Apkarian and R. J. Adams, Advanced gain-scheduling techniques
for uncertain systems, IEEE Trans. Control Systems Technology, vol.
6, pp. 2132, January 1998.
[6] G. Sachs, Path-attitude decoupling and flying qualities implications in
hypersonic flight, Aerospace Science and Technology, vol. 2, pp.
4959, January 1998.
[7] P. Gahinet, Explicit controller formulas for LMI-based H


synthesis, Automatica, vol. 32, pp. 10071014, July 1996.
[8] M. A. Bolender and D. B. Doman, Nonlinear longitudinal dynamical
model of an air-breathing hypersonic vehicle, Journal of Spacecraft
and Rockets, vol. 44, pp. 374387, March-April 2007.
[9] J. T. Parker, A. Serrani, S. Yurkovich, M. A. Bolender, and D. B.
Doman, Control-oriented modeling of an air-breathing hypersonic
vehicle, Journal of Guidance, Control, and Dynamics, vol. 30, pp.
856869, May-June 2007.
[10] H. Buschek and A. J. Calise, Uncertainty modeling and fixed-order
controller design for a hypersonic vehicle model, Journal of Guidance,
Control, and Dynamics, vol. 20, pp. 4248, January-February 1997.
WeD2.3
284

You might also like