Sei sulla pagina 1di 19

52nd AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics and Materials Conference<BR> 19th AIAA 2011-1769

4 - 7 April 2011, Denver, Colorado

The Effect of Rate Limited Actuator Dynamics on the


Adaptive Control of a Nonlinear Aeroservoelastic (ASE) System

Robert Bruce Alstrom1, Goodarz Ahmadi2, Erik Bollt 3, Pier Marzocca 4

Clarkson University, Potsdam, New York 13699-5725

ABSTRACT

In this paper a nonlinear airfoil equipped with a trailing edge flap. The wing-
flap structure is coupled to a simplified rate limited flight control actuator. The
closed loop system will be examined for its aeroservoelastic response to a
reference free adaptive feedback controller with nonlinear actuator dynamics in
the loop.

INTRODUCTION

Aerodynamic flutter refers to a subject that has evolved since the beginning of manned flight.
Aeroelasticity continues to occupy a prime role in the current design of advanced aircraft,
missiles and launch vehicles. With the advent of modern flight control systems, the disciplines of
aeroelasticity and structural dynamics are linked together by the interaction between the
aeronautical structure and the flight control system and associated sensors; this is known as
aeroservoelasticity (ASE). Aeroservoelasticty exists primarily because of the excitation of the
flight vehicles structural modes can cause both oscillatory aerodynamic loads and flight control
system demands that result from the airframe mounted motion sensors; the aeroservoelastic
interaction comes from these secondary responses which induce further excitation of structural
modes. It is common practice to apply notch and low pass filters to the flight control system in
order to meet the stability and gain margins required to clear an aircraft for flight. But in doing
so one can incur undesirable phase lags at the rigid body frequencies. The secondary oscillatory
dynamics often present as Limit Cycle Oscillations (LCO). There is a distinction to be made
between the LCOs caused by ASE interaction and those caused principally by the aero-structure
itself. Hence it is instructive to understand the mechanism by which limit cycles and chaotic
motions are created. A limit cycle is a standing periodic oscillation that is characterized in the
phase plane as a single loop; in some instances a multi-frequency limit cycle may appear as in
the case of a chaotic system that has been transformed by an appropriate control force. For a
stable limit cycle, the rate of energy input from the freestream is equal to the energy dissipation
rate. A limit cycle is unstable when the system continuously receives more energy than it is able
to dissipate and it will grow in amplitude. Conversely, if energy is extracted from the system,
then the oscillation will decay. It is this mechanism that makes it possible to force an aeroelastic

1
Graduate Student, Department of Mechanical and Aeronautical Engineering, AIAA Senior Member
2
Professor, Department of Mechanical and Aeronautical Engineering
3
Professor, Department of Mathematics and Computer Science
4
Associate Professor, Department of Mechanical and Aeronautical Engineering, AIAA Senior Member

Copyright 2011 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved.
system to switch limit cycles. We will revisit the notion of an unstable limit cycle and limit cycle
switching later. For now we will introduce examples of limit cycle switching and suppression.

The first such example as pointed out by Dimitriadias and Cooper [1] was work performed by
Holden et al [2]. Specifically, during a series of wind tunnel tests on a flutter model of a tail
plane, Holden and his colleagues noticed that applying a certain excitation signal caused an
LCO; the re-application of the same signal a short time later produced a limit cycle of larger
amplitude. The manipulation of limit cycles has a lot to do with the energy state of the system.
There are devices which have the capability to make instantaneous changes of its mass, stiffness
or damping; such devices are termed state switchable dynamical systems. For example if a
switchable stiffness element is build into a vibration absorber, the change in stiffness causes a
change in the resonant frequencies of the system and thus retuning the system. One can design
a switching rule-control law that extracts energy from the system [3, 4]. Lee et al [5]
demonstrated that one can use continuously varying stiffness and damping elements (or
nonlinear energy sinks NES) coupled to a van der Pol oscillator. By adjusting the parameters of
the NES indicated in Figure 1, suppression of LCOs can be achieved.

Figure 1. System configuration with: (a) grounded NES; (b) ungrounded NES [5]
Now lets revisit the discussion on unstable limit cycles also referred to as unstable periodic
orbits (UPOs); for a system with sufficient complexity (i.e. multiple degrees of freedom, number
and type of nonlinearities), a significant number of limit cycles can exist in its phase plane. As
such the key is to force a system into a stable limit cycle knowing the location of the unstable
limit cycles that surround it. A very similar argument was made by Pyragas [6]. Specifically, the
stabilization of unstable periodic orbits of a chaotic system is achieved by applying a combined
feedback with the use of a specially designed external oscillator or by a delayed self controlling
(Figure 2) feedback force without the use of an external force. Both of these methods do not
require any prior analytical knowledge of the system. These methods make use of the fact that
there are an infinite number of unstable periodic orbits contained in a chaotic attractor. The
delayed-self controlling method was successfully demonstrated for an aeroelastic system by
Ramesh et al. [7].

2
Figure 2. Schematic of delayed-feedback control system [6].
The majority of closed-loop aeroservoelastic work does not include the actuator dynamics. If the
actuator dynamics is modeled, it is observed that the nonlinearity is either in the aero-structure or
the actuator but tend not to be cascaded or in series under closed-loop control [16]. The other
combination seen in the literature is the use of a model that includes the trailing edge flap
characterized by a nonlinear stiffness. Further, the closed-loop controllers developed for this type
of system tend also to consider one nonlinearity at a time (i.e. either rate limiting or saturation)
[17]; real flight control actuators have multiple nonlinearities [18]. The inclusion of the actuator
dynamics is more common place in flight control system development work. The actuator
dynamics are often linear transfer function relationships. Will the adaptive control signal still be
able to suppress the chaotic motion with the actuator dynamics included? The aim of this work is
to investigate the effects of simulated nonlinear actuator dynamics on the designed control
signals capability to provide suppression to the aeroelastic system. To facilitate this evaluation,
a 2D nonlinear wing with a trailing edge flap from the Texas A& M Aeroelasticity Group will be
used in [19].This model allows for coupling of the actuation system in series with the aeroelastic
plant (Figure 3) as the flap displacement input to the system as given by Eqn. 1

( ) (1)

EHA Dynamical
System

LPF
k

Figure 3. Block diagram of adaptive controller with actuator

3
ADAPTIVE CONTROLLER
Controlling chaotic behavior mainly deals with the stabilization of unstable periodic orbits. The
stabilization of a fixed point by classical methods requires knowledge of the location in the phase
space. For many complex systems, the locations of the fixed points are not known a priori; as
such, adaptive control techniques that are capable of locating unknown steady states are
desirable. A simple adaptive controller for stabilizing unknown steady states can be designed
using ordinary differential equations (ODEs). One such controller utilizes a first order ODE that
represents a low pass filter (LPF). The filtered DC output of the filter estimates the location of
the fixed point, such that the difference between the actual and the filtered signal can be used as
a control signal. This control signal is then scaled by a proportional gain. The structure of the
system is represented in Figure 4.

Figure 4. Block diagram of adaptive controller. LPF denotes low pass filter [11]
In general mathematical terms, when considering an autonomous dynamical system as the one in
Figure 4, a description by a set of ordinary differential equations can be cast as:

( ) (2)

where the vector defines the dynamical variables and p is the total control force. The
scalar variable ( ) ( ( )) is a function of the systems states; for the aeroelastic system,
y() could be anyone of the four states i.e. pitch, plunge, or any of their rates. Suppose that
, that is the system has an unstable fixed point at x* that satisfies ( ) . If the
steady state value y*=g(x*) corresponding to the fixed point were known, one could try to
stabilize it using classical proportional feedback control.
( ) (3)
Suppose now, that the reference value y* is unknown. The objective will be to construct a
reference-free feedback perturbation that automatically locates and stabilizes the fixed point.
When the controller locates the fixed point, the control input should vanish i.e. no control power
should be dissipated in the closed-loop condition. The controller that satisfies the requirements
can be constructed from an ODE that represents a low pass filter given by the following
equation:

( ) (4)

Here, w is a controller variable and the parameter, represents the cut-off frequency of the
filter. The output of the filter provides an averaged input variable y(t). If y(t) oscillates about the
steady state value of y* one can expect that the output variable w(t) will converge to this value.
4
As result the reference value y* can be replaced with the output variable of the filter; that is the
control force can now be expressed in the following way:
( ) (5)
The complete control-loop is in fact a high-pass filter, since the second term in Eq. (8) is
obtained from the difference of the actual output signal and filtered by the LPF. The control
signal is proportional to the derivative of the controller variable w i.e.

( ) ( ) (6)

As , from Eq. (5), it follows that ( ) ( ); for large , the control signal becomes
proportional to the derivative of the output . When this happens, the controller behaves as a
simple derivative controller.

( ) ( ( ) ( )) (7)

For our numerical study, is set to zero so that the system is reference free and ( ) ( ).

THE NONLINEAR AEROELASTIC TEST APPARATUS (NATA)

The Aeroelasticity Group at Texas A&M University has conducted a number of experiments
using the apparatus shown below (Figure 5).The experiments performed have provided the
validation for the theoretical model presented in this report. The NATA test bed has been used
to study both linear and nonlinear aeroelastic behaviour as well as the development of control
laws for flutter suppression. The system consists of a NACA 0012 and is controlled by a full
span trailing edge control surface located at 20% chord. The pitch and plunge stiffnesses of the
NATA are provided by springs attached to cams with profiles designed to illicit a specific
response. For example, a parabolic pitch cam yields a spring hardening response. This is the
mechanism that causes the NATA to exhibit the limit cycle behaviour.

Figure 5. Aeroelastic System

5
The flexible mounting of the wing structure allows the NATA to move in two degrees of
freedom displacement, i.e. torsion and bending displacements. Bending or plunge is denoted by h
and torsion or pitch is denoted by . The derivation of the following equation can be found in
Fung. The governing equations of motion are given by:

h L
M h C h K (8)
M

In Equation 8, L and M are the aerodynamic lift and moment respectively. A quasi-steady
aerodynamic assumption is used in this model:

h
L U 2bcl 0.5 a b U 2bcl (9)
U U
h
M U 2b 2cm 0.5 a b U 2b 2cm
U U

c c c c
where l and l are the lift and moment coefficients per angle of attack, and m and m are

the lift and moment coefficients per control surface deflection. The nonlinear torsional stiffness
is obtained through experimental data and is given by:


k ( ) 2.82 1 22.1 1315.5 2 8580 3 17289 4 (10)

By combining equations 8 and 9, we can define the matrices presented in equation 8. They are
defined as follows:
m mx b ch Ubcl Ub2cl 0.5 a
M , C
I Ub cm c Ub3cm 0.5 a
2
mx b

kh U 2bcl bcl
K , Qf 2
0 U 2b2cm k b cm

Equation 8 is used to build the nonlinear aeroelastic model in SIMULINK (Figure 6). Note that
in Figure 6 there is a block called Nonlinear Stiffness Matrix; the structural nonlinearity has been
implemented using the function block. It allows one to read the angle of attack such that equation
8 can be implemented in SIMULINK. Figure 7 demonstrates the open loop response for a shear
center location of a= -0.75 and tunnel velocity of 25 m/s.

6
Figure 6. Nonlinear Aeroelastic SIMULINK Model

Figure 7. Open Loop Response (a=-0.75, U=25 m/s)

7
NONLINEAR ACTUATION SYSTEM MODELING & CLOSED LOOP SIMULATION

Todays aircraft employ 2 types of actuators: the electro-hydraulic (EHA-1) and the electro-
hydrostatic actuators (EHA-2). Both actuator types are typically used in aircraft flight control
and Stability Augmentation Systems (SAS) [19, 20]. A traditional aircraft servo-hydraulic
system has the components shown in Figure 8. The non-hydrostatic actuator (EHA-1) is
composed of four main components, the actuator control or command system, then main valve
actuation, the main valve and the ram.

Figure 8. Typical Electro-Hydraulic Actuator [18]

The electro-hydrostatic actuator (Figure 9) differs from the hydraulic version in that it does not
require a central hydraulic power supply which eliminates the need for plumbing. Instead the
EHA (hydrostatic) uses electrical cables. Specifically, the hydrostatic actuator is a Power-By-
Wire (PBW) actuator that utilizes the hydraulic pump to transfer the rotational motion of the
electrical motor to the actuator output. The EHA-2 is based principally on closed circuit
hydrostatic transmission; as a result the need for oil reservoirs and electro-hydraulic servo-valves
are eliminated [23]. A version of the electro-hydrostatic actuator designed by Moog Inc is
currently being flown on the F-35 Lightning II (Joint Strike Fighter).

Figure 9. Electro-Hydrostatic Actuator [23]

Due to the complexity of the nonlinear models for both types of EHA, we will make use of a
nonlinear first order rate limited actuator model for this analysis. This model still retains most of
the characteristics of interest including rate limited operations and actuator nonlinearities. The
selected model is realistic and it was used successfully in the study of Pilot Induced Oscillations
(PIO) [25] in the simulation and flight test of the NF-16D Variable Stability In-flight Simulator
Test Aircraft (VISTA).Often the same flight control system actuators allocated to the automatic

8
flight control system (AFCS) are also employed in flutter suppression and hence the same
actuator nonlinearities that cause flying qualities issues also cause aeroservoelastic interactions.
Initially it was thought that an actuator model with multiple nonlinearities would be effective in
this investigation but after several simulations it proved difficult to isolate the cause of the
closed-loop instability. As such the nonlinear first-order rate limited actuator model provides a
means for simple parametric investigation and demonstration of the research problem at hand.
Common actuator nonlinearities include saturation, friction, dead zones (or free play) and
hysteresis and rate saturations. In this study we will focus on the rate saturation nonlinearity. The
simplified model of the rated limited actuator is presented in (Figure 10)

Figure 10 Rate Limited Actuator [25]

Figure 11 Effective Nonlinear Gain for Open Loop Saturation Nonlinearity [25]

The saturation element for the configuration seen in Figure 11 is given by the describing function
[24]:

( ) [ ( ) ( )] (15)

By using a series expansion, keeping only the first order linear terms, the describing function can
be written as follows:
( ) (16)

9
By replacing Equation 16 can be cast as ( ) . The derivation in Slotine and
Li [24] employs four conditions that must be satisfied, one of which is that the properties of the
nonlinear elements do not vary with time; while this true, in our application, the signals entering
the actuator will be non-stationary in nature. The parameters of this model are defined in Figure
11.

When the error signal , the saturation point, a rate limited actuator induces a phase lag in
the closed loop system. This can cause limit cycles. From a frequency domain standpoint,
actuators have a flat bandwidth. When the amplitude is increased such that , the right
most side of magnitude and phase plots start to migrate toward the left hand side of the diagram;
physical this means that the actuator is operating with a decreased bandwidth and the output
signal will be clipped. For a more in-depth treatment of the effect of rate limiting, the reader is
directed to Klyde [25]. The actuator model is placed in series with the controller and comes
before the aeroelastic system in the given SIMULINK block diagram (Figure 12). Specifically,
the magenta colored blocks are the actuator blocks and the yellow blocks represent the
aeroelastic system and its outputs. The blue blocks are the components of the controller and the
red block is a switch that is used to engage the controller at different times in the simulation.

Figure 12 Closed Loop SIMULINK Model

For the purpose of this study, an airspeed of 25 m/s (above flutter) and a shear center is
coincident with the mean aerodynamic center have been selected; we will also confine our
analysis to sea level conditions. Before we can begin to investigate the issue of suppression with
actuator dynamics in the loop, we must first understand how the frequencies change with
airspeed and shear center location. Lets now introduce the structural coupling design procedure
used in this work, after which it will become clear how the velocity-frequency plot (Figure 13) is
used.

10
Figure 13 Velocity vs. Frequency

After some exploratory simulation runs, the following procedure was developed to find the
minimum feedback gain with an actuator in the loop.

1. Select the feedback gain, k without the actuator dynamics. This is done by using the
manual switch (See in Figure 13). Vary k until the angle-of-attack is approximately zero.
The gain that corresponds to a zero angle-of-attack is the first estimate of the feedback
gain.
2. Switch the simulation path to the actuator path and input the value of k obtained in Step
1. In this step, the frequencies for the actuator and the low pass filter as it is given in the
Section Adaptive Control must be selected. First the low pass filter cut-off frequency
must be chosen such that it attenuates the first mode; the first mode from the green curve
that represents the wing configuration of a=-0.75, the lowest velocity, 5 m/s where the
frequency, (or 10.83 rad/s) was selected; the cutoff frequency, is set
equal to 10 rad/s, less that than the first mode at 5 m/s. Next the actuator frequency gain,
should be chosen such that there is sufficient frequency separation between the first
mode frequency at 40 m/s (9.765 Hz or 61.35 rad/s); graphically this means that if one
was to scribe a straight line at some desired actuator frequency, it should not intersect the
velocity frequency boundary of interest. The actuator frequency is then set equal to
30 Hz (or 188.49 rad/s). To summarize we have now selected the actuator and low pass
filter frequencies and the estimated feedback gain, k from Step 1.
3. Vary the rate limit, until the angle-of-attack, is approximately zero. The rate limit
obtained in this step is the critical value of the rate limit i.e. the final rate limit must
selected such that ( ) ( )
4. Using the or some value greater prescribed by the inequality in Step 3, return to the
feedback gain, k and vary the parameter until the angle-of-attack is approximately zero.
The frequency and rate limit parameters should not be changed. The new value of the

11
feedback gain obtained in this step is the value of feedback gain required to ensure stable
limit cycle and flutter suppression with the actuator dynamics in the loop.

RESULTS AND DISCUSSION

The closed loop bifurcation without actuator dynamics (in black-Figure 14) shows that a
feedback gain of 80 N/m will give zero angle-of-attack. Compare this result with the red curve; it
appears to have a lower and upper branch and a discontinuity. The significance of these features
is best understood by studying Figure 15. The discontinuities in Figure 15 can either be
hysteretic or regions of intermittent switching indicating a mode transition. For this work, when
the gain is selected such that the feedback gain is greater than 44 N/m, the closed loop system
exhibits intermittent switch and presents as chaotic motion as seen in the phase plane (Figure
18). The mode transition can be visualized by the Lissajous curve.A Lissajous figures indicate
the periodicity of the lower branch and the less steady dynamics of the upper branch as well as
phase changes. Figure 16 shows that the actuator input to the aeroelastic system locks in and
confines the wing to a limit cycle. Figure 17 is indicates rate limiting by the wave like closed
trajectories which translates to chaotic motions (again see Figure 18). The implications for the
control system design lies in the fact that for this set of model parameters, the flat region of the
lower branch is smaller; that is the designer has a smaller range of feedback gains that will
produce zero displacements is less

Figure 14 Controller Bifurcation Diagram

12
Figure 15. Amplitude Response Curve [28]

Figure 16. Lissajous Curve (Lower Branch)

13
Figure 17. Lissajous Curve (in discontinuity)

Figure 18 Closed Loop Response in hysteresis region

14
Figure 19 Closed Loop actuator dynamics

Figure 20 Controller Bifurcation Diagram (reduced rate limit)

15
Figure 21 Controller Bifurcation Diagram (increased rate limit)

Actuator rate limiting occurs when the input rate to the control surface exceeds the hydraulic
and/or mechanical capability of the control actuator. One tell that the actuator is rate limited in
Figure 19 by looking at the output, . The output exhibits periodicity and has a saw tooth profile.
The actuator rate is a clipped signal. If the rate limit, were reduced by 50%, it can be
observed that the hysteresis zone moves to the left. In addition, the length of the flat portion of
the stability boundary decreases. The flat portion is important because it is the set of feedback
gains that yield a zero angle-of-attack when the loop is closed on the actuator. Conversely, when
the rate limit is increased by a factor of 2 the flat portion of the stability boundary increases,
indicating that the hysteresis/rate limit zone moves to the right (Figure 21).

FUTURE WORK

NASA has initiated an Integrated Resilient Aircraft Control (IRAC) initiative under the Aviation
Safety Program. The main thrust of this initiative is to advance the state-of-the-art technology in
order to facilitate a design option that allows for increased resiliency to failures, damage, and
upset conditions. These adaptive flight control systems will have the capability to automatically
adjust the control feedback and command paths to regain stability in the closed loop
configuration. One of the consequences of changing the control feedback and command path
configuration is the occurrence of aeroservoelastic (ASE) interaction which results in undesirable
limit cycle oscillations. The combination of changing structural behavior with changing control
system gains results in a system with a probability of adverse interactions that is very difficult to
predict a priori. Onboard, measurement based methods are required to ensure that the system
adjusts to attenuate any adverse ASE interaction before a structural system can become entrained

16
in sustained limit cycle and vehicle damage occurs. This system must work in concert with the
adaptive control system to restore nominal rigid body performance as much as possible without
exacerbating the situation with ASE interactions. To that end Li [26] developed an in-flight
narrow bandpass filter (NBWF) detection method that is coupled with an adaptive notch filter
that was inserted into the command path to attenuate limit cycle oscillations in the vehicle flight
dynamics. In this study we have seen that the rate limiting nonlinearity can induce sever closed
loop instability. Originally, the RLPF was employed in the investigation into pilot induced
oscillations (PIOs) and there prevention during the aggressive, high gain maneuvers performed
by fighter aircraft [27]. PIOs can result in departure from controlled flight. The RLPF was also
inserted into the command path. Since the same flight control actuator nonlinearities that can
cause PIOs can also induce structural coupling and limit cycle oscillations, it may be of interest
to explore whether the RLPF can attenuate in flight limit cycle oscillation events.

Figure 21. Nonlinear Rate Limit Pre-filter (RLPF) [27]

CONCLUSIONS

In this study we have successfully illustrated that that nonlinear actuator dynamics can
destabilize a closed loop system. We also show that such systems can be stabilized without any
reduction in actuator bandwidth. Regions of hysteresis can be very dangerous in the sense that a
closed loop stable system can become chaotic and may lead to catastrophic failure. Also, the
hysteresis region may limit the use of a full adaptive controller. In a simple way, this study
demonstrates that the actuator can be designed to attenuate structural modes without the use of a
notch filter; this is can probably attributed to the fact that in this evaluation the actuator was
tuned such that there was no structural coupling, this is not always the case in major aircraft
development programs. The final parameter configuration based on the design method presented
in this work is as follows: The low pass filter and actuator frequencies are 10 rad/s and 188.5
rad/s respectively. The feedback gain (with nonlinear actuator dynamics) is 40 to 44 N/m with
the bending displacement as the feedback signal. For the shear center location of a=-0.75 at
U=25 m/s, this represents a 50% decrease in the control power required to maintain closed loop
stability when actuator dynamics are included. The final rate limit for the actuator is 15 deg/s.

17
REFERENCES

1. Dimitriadis, G., Copper, J.E. Limit Cycle Oscillation Control and Suppression, The
Aeronautical Journal, Vol. No. 1023, May 1999, pp 257-263
2. Holden, M., Brazier, R.E.J. and Cal, A.A. Effects of structural nonlinearities on a
tailplane flutter model, IFASD, Manchester UK, 1995
3. Cunefare, Kenneth, A., De Rosa, Sergio, Sadegh, Nader and Larson, Gregg, State
Switched Absorber for Semi-Active Structural Control, Journal of Intelligent Material
Systems and Structures, Vol. 11, April 2000, pp 300-310
4. Cunefare, Kenneth, A., State-Switched Absorber for Vibration control of Point-
Excited Beams, Journal of Intelligent Material Systems and Structures, Vol. 13, March
2002, pp 97-105
5. Lee, Young S., Vakakis, Alexander F., Bergman, Lawrence A., and McFarland,
Michael D., Suppression of limit cycle oscillations in the van der Pol oscillator by
means of passive nonlinear energy sinks, Structural Control and Health Monitoring,
Vol. 13, 2006 pp 41-75.
6. Pyragas, K., Continuous control of chaos by self-controlling feedback, Physics Letters
A 170 (1992), pp 421-428
7. Ramesh, M., Narayanan, S., Controlling Chaotic Motions in a Two-Dimensional
Airfoil Using Time-Delayed Feedback, Journal of sound and Vibration (2001) 239(5),
pp 1037-1049.
8. Zhao, Y.H., Stability of a two-dimensional airfoil with time-delayed feedback control
Journal of Fluids and Structures, Vol. 25, 2009, pp 1-25
9. Marzocca, P., Librescu, L., Silva, W.A., Time-delay effects on Linear/Nonlinear
Feedback Control of Simple Aeroelastic Systems, Journal of Guidance, Control and
Dynamics, Vol. 28, No. 1, January-February 2005.
10. Rubillo, C., Marzocca, P., Bollt, E., Active Aeroelastic Control of Lifting Surfaces via
Jet Reaction Limiter Control International Journal of Bifurcation and Chaos, Vol. 16,
No. 9, 2006, pp 2559-2574
11. Pyragas, K., Pyragas, V., Kiss, I.Z., and Hudson, J.L., Adaptive Control of Unknown
Steady States of Dynamical Systems, Physical Review E 70, 026215 (12 pages), 2004
12. Yuan, Y., Yu, P., Librescu, L. and Marzocca, P., Aeroelasticity of Time-Delayed
Feedback Control of Two-Dimensional Supersonic Lifting Surfaces, Journal of
Guidance, Control, and Dynamics, Vol. 27, No. 5, 2004, pp. 795-803.
13. Librescu, L., Marzocca, P., Silva, W.A., Aeroelasticity of 2-D lifting surfaces with
time-delayed feedback control, Journal of Fluids and Structures, Vol. 20, No. 2, 2005,
pp. 197-215
14. Yuan, Y., Yu, P., Librescu, L. and Marzocca, P., Implications of time-delayed feedback
control on limit cycle oscillation of a two-dimensional supersonic lifting surface,
Journal of Sound and Vibration, Vol. 304, No. 3-5, 2007, pp. 974-986
15. Alstrom, Robert Bruce, Marzocca, Pier, Bollt, Erik and Ahmadi, Goodarz, Controlling
Chaotic Motions of a Nonlinear Aeroelastic System Using Adaptive Control Augmented
with Time Delay, AIAA GNC/AFM/MST/ASC/ASE 2010 Conference. Toronto
,Ontario Canada, August 2010.

18
16. Dimitriadis, G., Cooper, J.E., Characterization of the behavior of a Simple
Aeroservoelastic System with Control Nonlinearities, Journal of Fluids and Structures,
Vol. 14, No. 8, 2000, pp 1173-1193
17. Demenkov, Max, Goman, Mikhail, Suppressing Aeroelastic Vibration via Stability
Region Maximization and Continuation Techniques, UKACC Control Conference,
Manchester England, 2008. http://www.control2008.org/
18. Taylor, Richard, Pratt, Roger W., and Caldwell, Brian D., Effect of Actuator
Nonlinearities on Aeroservoelasticity, Journal of Guidance, Control and Dynamics,
Vol. 19, No. 2, March-April 1996.
19. Ko, J., Kurdilla, A.J., and Strganac, T.W., Nonlinear Control of a prototypical wing
section with torsional nonlinearity, Journal of Guidance, Control and Dynamics, Vol.
20, No. 6, pp 1181-1189.
20. Edwards, John W., Analysis of an Electrohydraulic Aircraft Control-Surface Servo and
Comparison with Test Results, NASA TN D-6928, 1972
21. Fielding, C., Flux, P.K., Non-linearities in flight control systems, The Aeronautical
Journal, No. 2838, November 2003.
22. Stirling, R., Actuation system jump resonance characteristics University of Bristol,
Department of Engineering Report RS/2/84 (1984).
23. Kang, Rongjie, Mare, Jean Charles, and Jiao, Zongxia, Nonlinear Modeling and
Control Design of Electro-Hydrostatic Actuator, Proceedings of the 7th JFPS
International Symposium on Fluid Power, TOYAMA 2008, Sept 15-18, 2008.
24. Slotine, Jean-JacquesE., Weiping Li. Applied Nonlinear Control. Upper Saddle River
NJ: Prentice Hall, 1991.
25. Kylde, David H, Unified Pilot-Induced Oscillation Theory, Volume 1:PIO Analysis
With Linear and Nonlinear Effective Vehicle Characteristics, Including Rate Limiting.
WL-TR-96-3028. AFRL, Wright Patterson AFB OH, December 1995.
25. Li, Xiaohong, L., Brenner, Martin J., Practical Aeroservoelasticity In-Flight
Identification and Adaptive Control, AIAA GNC/AFM/MST/ASE 2010 Toronto
Ontario, Canada.
26. Chapa, Michael J., A Nonlinear Pre-Filter to Prevent Departure and/or Pilot Induced
Oscillations (PIO) due to Actuator Rate Limiting.
27. Khalak, A, Williamson, C.H.K., Motions, Forces and Mode Transitions in Vortex-
Induced Vibrations at Low Mass-Damping, Journal of Fluids & Structures (1999) 13,
813-851.

19

Potrebbero piacerti anche