You are on page 1of 19

JOURNAL OF GUIDANCE, CONTROL, AND DYNAMICS

Vol. 28, No. 5, September–October 2005

Robust Integrated Flight Control Design Under Failures,


Damage, and State-Dependent Disturbances

Jovan D. Bošković,∗ Sarah E. Bergstrom,† and Raman K. Mehra‡


Scientific Systems Company, Inc., Woburn, Massachusetts 01801

A robust integrated fault-tolerant flight control system is presented that accommodates different types of ac-
tuator failures and control effector damage, even while rejecting state-dependent disturbances. It is shown that
a decentralized failure detection, identification, and reconfiguration system, combined judiciously with adaptive
laws for damage estimates and variable structure adjustment laws for disturbance estimates, yields a stable system
despite simultaneous presence of failures, damage and disturbances. The proposed system is well suited for the case
of first-order actuator dynamics. The properties of the proposed algorithms are illustrated on a medium-fidelity
Downloaded by INDIAN INSTITUTE OF TECHNOLOGY on April 3, 2019 | http://arc.aiaa.org | DOI: 10.2514/1.11272

nonlinear simulation of Boeing’s Tailless Advanced Fighter Aircraft.

I. Introduction damage FDIR subsystems with disturbance rejection algorithms,


which results in a stable overall system.
T O increase the autonomy of unmanned aerial vehicles (UAV)
and the duration of their autonomously performed missions,
the flight control systems (FCS) for UAVs need to be capable of
Most of the available FDIR techniques for flight-critical faults and
failures address either actuator failures or control effector damage.
accommodating a large class of subsystem and component failures An integration of the resulting FDIR subsystems and minimiza-
and structural damage without substantially affecting the perfor- tion of their interactions has not been studied. In addition, some
mance of the overall system. In this context, important components of the available FDIR algorithms are well suited for linear models,
of the FCS are on-line failure detection, identification, and recon- whereas those suitable for nonlinear aircraft dynamics are based
figuration (FDIR) algorithms the role of which is to rapidly and on-neural networks where a large number of parameters needs to be
accurately detect and identify different failures and faults and ap- adjusted. In addition, these algorithms are based on direct adaptive
propriately reconfigure the control laws to maintain the stability of control and, hence, do not provide FDIR information that is use-
the system. ful for condition-based maintenance. An overview of the existing
To be able to accommodate different faults and failures, the FDIR results is given next.
system commonly needs to effectively integrate several FDIR algo- Several results involving FDIR based on linearized models of
rithms. For instance, sensor failure algorithms need to be integrated aircraft dynamics have been reported.1−10 An application of multi-
with actuator failure accommodation subsystem, and the resulting variable adaptive control techniques to flight control reconfiguration
system needs to be integrated with the algorithms capable of accom- was considered.1 The objective was to redesign automatically flight
modating structural damage and resulting disturbances. One of the control laws to compensate for actuator failures or surface damage.
major issues in the FDIR design is the complexity of the resulting A reconfigurable control scheme based on on-line parameter esti-
algorithms and their interactions that, if not taken into account, can mation and linear quadratic control algorithms was evaluated on
lead to substantial performance deterioration and even instability in a nonlinear simulation of the F-16 aircraft.2 A scheme combining
the system. The main challenge in this context is to arrive at a FDIR constrained least-squares identification of uncertain stability and
system that is capable of accommodating such faults and failures control derivatives and a receding horizon optimal control law was
even while minimizing the interactions between the FDIR subsys- developed and flight tested on a VISTA/F-16 aircraft.3 A similar
tems. In this paper a design of an integrated algorithm that achieves approach based on static constrained identification and prior infor-
the FDIR objective and maintains the desired performance of the mation was used4 to estimate time-varying parameters arising as a
system is described. result of failures and was evaluated on a F-16 aircraft simulation. A
The focus of the study presented here is on flight-critical failures common feature of these schemes is that they were developed for
and faults. In particular, the analysis and design are concentrated on linearized models of aircraft dynamics and are well suited for single
actuator failures and control effector damage that generates large actuator failures. The techniques are based on on-line estimation
state-dependent disturbances. Both types of faults, if not rapidly of a large number of uncertain parameters associated with failures,
accommodated, can lead to substantial performance degradation and which can result in slow response and large transients.
instability of the closed-loop system. The design challenge in this A neural-network-based adaptive control approach5 was devel-
context is an integration of the actuator failure and control effector oped and evaluated on a simulation of F/A-18 aircraft. A similar
approach was applied to the Boeing’s Tailless Advanced Fighter Air-
craft (TAFA).6 The neural network was used to compensate for the
Presented as Paper 2003-5490 at the AIAA Guidance, Navigation, and plant inversion error as a result of modeling uncertainties, failures,
Control Conference, Austin, TX, 11–14 August 2003; received 3 June 2004; and damage. This algorithm was combined with on-line control allo-
revision received 19 October 2004; accepted for publication 7 December cation based on identification of control derivatives. The properties
2004. Copyright  c 2005 by Scientific Systems Company, Inc.. Published of the overall control system were evaluated on a high-fidelity TAFA
by the American Institute of Aeronautics and Astronautics, Inc., with per- simulation.6 These algorithms were successfully evaluated through
mission. Copies of this paper may be made for personal or internal use, hardware-in-the-loop simulations and flight testing on a X-36 tail-
on condition that the copier pay the $10.00 per-copy fee to the Copyright less fighter aircraft.7 Although these algorithms have the potential to
Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923; include deal simultaneously with actuator failures and nonlinearities arising
the code 0731-5090/05 $10.00 in correspondence with the CCC.
∗ Autonomous and Intelligent Control Systems Group Leader, 500 W. as a result of control effector damage, their utility in that context has
Cummings Park, Suite 3000; jovan@ssci.com. Senior Member AIAA. yet to be demonstrated. In addition, neural networks require on-line
† Research Engineer, 500 W. Cummings Park, Suite 3000; jovan@ssci. adjustment of a large number of parameters (weights), which can
com. require very complex tuning in different flight regimes.
‡ President and CEO, 500 W. Cummings Park, Suite 3000; rkm@ssci.com. It is of interest to extend the existing results in the area of FDIR
Member AIAA. to the case of simultaneous actuator failures and control effector
902
BOŠKOVIĆ, BERGSTROM, AND MEHRA 903

damage accommodation, arrive at a design that adjusts a moderate II. Problem Statement
number of adaptive parameters, minimizes the interactions between In this paper the focus is on the class of models of aircraft dy-
the resulting FDIR subsystems, and effectively compensates for the namics of the form
nonlinearities arising as a result of the damage. Some of the results
that address some of the aspects of this problem are described next. ż 1 = f¯(z) (1)
The approach from Ref. 8 addresses the problem of a large number
of adjustable parameters encountered in the existing FDIR designs ż 2 = f 0 (z) + G 0 (z)u + ξ0 (z) (2)
by parameterizing a large class of actuator failures using only two
parameters, and, based on on-line estimation of these parameters u̇ = −λ(u − u c ) (3)
and their use in the reconfigurable control law, ensures fast failure +
where z : IR → IR denotes the state vector; z 1 : IR → IR
n + (n − p)
,
FDI and robust control reconfiguration. In another paper,9 control z 2 : IR+ → IR p , z = [z 1T z 2T ]T , f¯ : IRn → IR(n − p) , f 0 : IRn → IR p ,
effector damage problem was addressed using both gradient adap- G 0 : IRn → IR p × m , u c : IR+ → IRm , and u : IR+ → IRm denote re-
tation and a variable structure algorithm. In both cases the stability spectively the controller output vector and the plant input vector;
of the overall system is ensured despite damage of multiple control ξ0 : IRn → IR p denotes an uncertain disturbance vector; and λ  1
effectors. However, it was also assumed that there are no damage- denotes the actuator gain. It is assumed that the preceding model
generated nonlinearities. describes the dominant dynamics of the aircraft. In such a case, the
To address the issues arising in the context of integrated FDIR, state variables include total velocity V ; angle of attack α; sideslip
in this paper a robust integrated fault-tolerant flight control design angle β; angular velocities p, q, and r ; and attitude angles φ, θ ,
Downloaded by INDIAN INSTITUTE OF TECHNOLOGY on April 3, 2019 | http://arc.aiaa.org | DOI: 10.2514/1.11272

that effectively compensates for actuator failures, control effector and ψ. In addition, it is assumed that the nonlinearities f 0 (z) and
damage, and large damage-generated state-dependent disturbances G 0 (z) are associated with nominal aircraft dynamics [i.e., the dy-
is proposed. Advantages of the proposed technique are as follows: namics in flight regimes in which there are no failures or damage and
1) It solves the FDIR problem for simultaneous actuator failure where ξ0 (z) = 0], and ξ0 (z) is an unknown nonlinearity generated
and control effector damage even while rejecting resulting large by failures or damage.
state-dependent disturbances and minimizing interactions between The preceding model is subject to the following assumption.
the FDIR subsystems. Assumption 1:
2) It uses a moderate number of adjustable parameters (3m + 6, a) State of the system is measurable.
where m is the number of control effectors). This number can be b) For a closed bounded set of states Sz , G 0 (z)G 0T (z) is invertible
further decreased if the FDI for actuator and control effectors that for all z ∈ Sz .
are less used in a particular flight regime is turned off. c) f 0 (z), G 0 (z), and ξ0 (z) are sufficiently smooth functions (func-
3) It is well suited for nonlinear aircraft dynamics. As shown tionals) of their argument.
in the paper, the proposed approach is well suited for nonlinear d) m > p.
models that are affine in the control input and are characterized by e) At least p actuators are highly reliable and can be assumed to
sufficiently smooth nonlinearities. be fault tolerant over an interval of interest.
4) It extends previous results by the authors,8 developed for lin- f) |ξ0i | ≤ c0i + d0i ϕ0i (z), where ϕ0 (z) = [ϕ01 (z) ϕ02 (z) . . .
earized models of aircraft dynamics and actuator FDIR only, to ϕ0 p (z)]T is a known vector function of the state, ϕ0 (0) = 0, and
nonlinear aircraft dynamics and simultaneous actuator failure and c0i ≥ 0 and d0i ≥ 0 are known.
control effector damage accommodation. Because the nonlinear model (1) and (2) describes the dynamics of
The proposed approach is based on the decentralized FDIR the dominant state variables, these are commonly measurable in the
scheme for actuator failure accommodation8 and is extended here to case of advanced fighter aircraft. Hence, assumption 1a is justified
include control effector damage and nonlinear aircraft dynamics and in most practical situations. In the case of gain-scheduled models
state-dependent disturbances. The main idea behind the proposed of aircraft dynamics, the aerodynamics effects are modeled as A(z)
approach is described next. The overall FDIR system includes de- and B(z), where the matrices A and B depend on several variables
centralized FDI systems for each of the actuators and a damage including the Mach number, altitude, angle of attack, and sideslip
and disturbance estimation module that estimates the extent of the angle. Each element of A and B can then be obtained from the
control effector damage and resulting state-dependent disturbances. look-up tables and represented as a multidimensional polynomial
The system also includes a robust reconfigurable controller that uses of these state variables. In most of the flight regimes, the resulting
the estimates of failure- and damage-related parameters and distur- matrix B(z) is such that B(z)B T (z) is invertible over a domain.
bances to ensure robust tracking and is shown in Fig. 1. In the case Assumption 1b is justified in such cases. Although assumption 1c is
of failure or damage, the FDIR system will accurately detect the commonly justified in the case of nominal nonlinear aircraft dynam-
failed actuator or control effector if there is enough persistent exci- ics (i.e., the case when ξ0 = 0), it might not be satisfied in the case
tation in the system. However,even in the case of poor excitation of of the nonlinearity ξ0 (z) generated by failures or damage. How-
the signals, the overall system will be stabilized despite the simul- ever, as a first approximation, smoothness of ξ0 (z) is assumed to
taneous effect of failures, damage, and disturbances. In such a case keep the problem analytically tractable. Assumption 1d is justified
additional tests can be run to accurately identify the failure. The in the case of advanced fighter aircraft characterized by a high level
design of the system from Fig. 1 is described in detail in this paper. of control effector redundancy. Because of assumption 1e, up to
m − p multiple simultaneous failures are allowed to occur. This is
not a restrictive assumption for modern combat aircraft because, for
instance, in the case of Boeing’s TAFA, m − p = 6. It is also noted
that some previous information regarding the nonlinearity ξ0 (z) is
assumed. This is justified in the case when the effect of failures
and damage on the aerodynamics is explicitly modeled based on
wind-tunnel data, similarly as in the case of wing damage.11
The model (1) and (2) is not in a convenient form because model-
reference control design results in a nonlinear reference model that
also depends on the state of the system.12 On the other hand, it turns
out that the models of aircraft dynamics have some convenient prop-
erties that can be used to transform them to a controllable canonical
form. For this reason the following assumption is considered:
Assumption 2: Vector f¯(z) can be expressed in the form
Fig. 1 Proposed robust supervisory FDIR system (© 1999–2004 Scien- f¯(z) = F(z 1 )z 2 , where F(z 1 ) is smooth and invertible on a closed
tific Systems Company, Inc.). bounded domain Sz .
904 BOŠKOVIĆ, BERGSTROM, AND MEHRA

This assumption is justified in the case when z 1 = [φ θ ψ]T , and where W = W T > 0 denotes a control allocation matrix. Upon sub-
a subvector z̄ 2 of vector z 2 is the vector of angular velocities in the stituting expression (14) into Eq. (2), one obtains
body frame z̄ 2 = [ p q r ]T . In such a case one has
ẋ1 = x2 , ẋ2 = Am x + Bm r
    
φ̇ 1 sin φ tan θ cos φ tan θ p The reference model equations (12) and (13) are subtracted from
 θ̇  = 0 cos φ − sin φ   q (4) the preceding system next to obtain
ψ̇ 0 sin φ sec θ cos φ sec θ r
ė1 = e2 , ė2 = Am e
= R(φ, θ)ω (5) Because the matrix A0 = [[0 I ]T AmT ]T is asymptotically stable,
it follows that limt → ∞ e(t) = 0; hence, the control objective is
where ω = [ p q r ]T . It can be readily shown that R(φ, θ) is invert- achieved in the case with no disturbances and actuator dynamics.
ible for all θ ∈ (−π/2, π/2). In the case with disturbances, an observer is first designed in the
Coordinate transformation: Under assumption 2, F(z 1 ) is now form
chosen as the transformation matrix for the z 1 subsystem, and a time
derivative of Eq. (1) is taken to obtain x̂˙ 1 = x̂2 (15)

z̈ 1 = Ḟ(z 1 )z 2 + F(z 1 )ż 2 x̂˙ 2 = Am (x̂ − x) + f (x) + G(x)u + ξ̂ (16)


 
Downloaded by INDIAN INSTITUTE OF TECHNOLOGY on April 3, 2019 | http://arc.aiaa.org | DOI: 10.2514/1.11272

∂F where ξ̂ denotes an estimate of ξ .


= F(z 1 )z 2 z 2 + F(z 1 )[ f 0 (z) + G 0 (z)u + ξ0 (z)] Because ξ is state dependent, standard adaptive control tech-
∂z 1
niques cannot be used to estimate it. Hence, the focus is on the
= f 1 (z 1 , z 2 ) + G 1 (z 1 , z 2 )u + ξ1 (z 1 , z 2 ) (6) variable-structure-control approach.13
Variable Structure Observer: Let ê = x̂ − x and êi = x̂i − xi ,
where f 1 (z 1 , z 2 ) = [(∂ F/∂z 1)F(z 1 )z 2]z 2 + F(z 1) f 0 (z), G 1 (z 1 , z 2) = i = 1, 2 denote the estimation errors. Upon subtracting the plant
F(z 1 )G 0 (z), and ξ1 (z 1 , z 2 ) = F(z 1 )ξ0 (z). The matrix G 1 (z) G 1 (z)T equation from expressions (15) and (16), one obtains
is invertible on the domain Sz because, under assumptions 1b and
2, both G 0 (z)G 0T (z) and F(z 1 ) are invertible on the same domain. ê˙ 1 = ê2 (17)
The following change of variables in now introduced: x1 = z 1 ,
and x2 = ż 1 . The relationship between z 2 and x2 is obtained from ė2 = Am ê + ξ̂ − ξ (18)
the inverse transformation: z 2 = F −1 (z 1 )ż 1 = F −1 (x1 )x2 . Hence, The preceding expression can be rewritten in a compact form as
ẋ1 = x2 (7)
ê˙ = A0 ê + C0 (ξ̂ − ξ ) (19)
 −1
  −1

ẋ2 = f 1 x1 , F (x1 )x2 + G 1 x1 , F (x1 )x2 u where A0 is defined earlier, and
   
+ξ1 x1 , F −1 (x1 )x2 (8) C0 =
0
Ip × p
The resulting model of the aircraft dynamics is of the form
Let a tentative Lyapunov function be of the form
ẋ1 = x2 (9)
V (ê) = 12 ê T P ê
ẋ2 = f (x) + G(x)u + ξ(x) (10)
where P = P T > 0 is a solution to the Lyapunov matrix equation
u̇ = −λ(u − u c ) (11) A0T P + P A0 = −Q, where Q = Q T > 0.
The first derivative of V along the motions of Eq. (19) yields
where f (x) = f 1 (x1 , F −1 (x1 )x2 ), G(x) = G 1 (x1 , F −1 (x1 )x2 ), and
ξ(x) = ξ1 [x1 , F −1 (x1 )x2 ]. The preceding model is now in a conve- V̇ (ê) = − 12 ê T Q ê + ê T P̄(ξ̂ − ξ )
nient form, and the design of a reference model is straightforward,
as shown next. Because the state of the original system (z 1 , z 2 ) is where P̄ = PC0 . Let P̄ = [ pi j ]. The preceding expression can be
measurable, the state of the transformed system is also measurable rewritten as
because x1 = z 1 and x2 = F(z 1 )z 2 . 1 p n
1
The desired dynamics of the aircraft in the transformed state space V̇ (ê) = − ê T Q ê + (ξ̂i − ξi ) pji ê j ≤ − ê T Q ê
2 i =1 j =1
2
is now chosen in the form

ẋm1 = xm2 (12)


p
n
p
n
+ [ci + di ϕi (x)] pji ê j + ξ̂i pji ê j
ẋm2 = Am xm + Bm r (13) i =1 j =1 i =1 j =1

The elements of the vector ξ̂ are now chosen as


where xm = [xm1 T
xm2 ]T denotes the state of the reference model;
+
xm : IR → IR , matrix A0 = [[0 I ]T AmT ]T is asymptotically sta-
m

n
ble; and r : IR+ → IR p denotes a vector of bounded piecewise con- ξ̂i = −[ci +di ϕi (x)] sign pji ê j , i = 1, 2, . . . , p (20)
tinuous reference inputs. j =1
Control objective 1: In the case with no failures, design a control
input u c (t) such that x(t) − xm (t) ≤ for all time despite the effect The derivative of V is now
of the disturbance ξ(x(t)). 1 p n
V̇ (ê) ≤ − ê T Q ê + [ci + di ϕi (x)] pji ê j
2 i =1 j =1
III. Disturbance Rejection Controller
A baseline controller is first designed for the case when actuator

p n
dynamics can be neglected (i.e., u = u c ) and with no disturbances 1
− [ci + di ϕi (x)] pji ê j ≤ − ê T Q ê < 0
(i.e., ξ ≡ 0): j =1 2
i =1

u c = W G T (x)[G(x)W G T (x)]−1 [− f (x) + Am x + Bm r ] (14) hence ê is bounded and limt → ∞ ê(t) = 0.


BOŠKOVIĆ, BERGSTROM, AND MEHRA 905

The controller (14) is next modified as


u c = W G T (x)[G(x)W G T (x)]−1 [− f (x) + Am x + Bm r − ξ̂ ] (21)
Upon substituting the control input (21) into Eq. (16), one obtains
x̂˙ 1 = x̂2 , ẋ2 = Am x̂ + Bm r
The reference model (12) and (13) is next subtracted from the pre-
ceding expressions to obtain
a) Lock-in-place c) Hard-over
ėm1 = em2 , ėm2 = Am em
where em = x̂ − xm and emi = x̂i − xmi , i = 1, 2. Hence limt → ∞
em (t) = 0. It is noted that em = ê + e, where e = x − xm , denotes
the tracking error, which implies that e = em − ê. Because ê(t) has
already been shown to tend to zero asymptotically, it follows that
limt → ∞ e(t) = 0, which ensures that the control objective is met.
Remark: The variable structure adaptive algorithm is robust to
arbitrary time variations in the vector ξ . Its main disadvantage is
Downloaded by INDIAN INSTITUTE OF TECHNOLOGY on April 3, 2019 | http://arc.aiaa.org | DOI: 10.2514/1.11272

that, because it contains a sign function, because of imperfections


of the switching devices, it results in chattering of the signals in the b) Float d) Loss of effectiveness
system. Hence such algorithms are commonly approximated either
by a saturation function or in the following way: Fig. 2 Types of control effector failures.

sign (η) ∼
= η/(|η| + δ) one has that u̇ i (t) = 0 for t ≥ tFi , and u(t) = u(tFi ) for all t ≥ tFi . In
where 0 < δ 1. If the sign function in the algorithm (20) is approx- the case of LIP, u(tFi ) has the value of u(tFi− ), whereas in the case of
imated in such a way, the resulting tracking error will be ultimately HOF it moves to the upper or lower position limit.
uniformly bounded (UUB) rather than tend to zero asymptotically,
and the size of the UUB set will depend on δ. B. Uncertainty
Uncertainty associated with each of the actuator models is caused
IV. Actuator Failures and Control Effector Damage by 1) unknown time of failure tFi , 2) unknown LOE coefficient ki ,
One of the important steps in the design of a FDIR system is failure and 3) unknown value at which the actuator locks.
and damage modeling and parameterization. In this section differ-
ent failure and damage cases are described, and the corresponding C. Control Effector Damage
mathematical models are given. The control objective under failures Control effector damage can be modeled using the diagonal con-
and damage is also stated. trol effector damage matrix D whose elements di are equal to one in
the no-failure case (i.e., D N = I ), while in the case of control effec-
A. Actuator Failures tor damage assume values over an interval [ di , 1], where di 1,
Typical actuator failures include 1) lock in place (LIP); 2) hard- and each value of d is proportional to the percentage of the loss of
over failure (HOF); 3) float; and 4) loss of effectiveness (LOE). In surface. The resulting model is of the form
the case of LIP failures, the actuator “freezes” at a certain condition
ẋ1 = x2 (23)
and does not respond to subsequent commands. HOF is character-
ized by the actuator moving to and remaining at the upper or lower ẋ2 = f (x) + G(x)Du + ξ(x) (24)
position limit regardless of the command. The speed of response
is limited by the actuator rate limit. Float failure occurs when the where D = diag[d1 d2 . . . dm ], and ξ(x) is a nonlinearity arising as
actuator contributes zero moment to the control authority. Loss of a result of the asymmetry of the damaged aircraft such that, in the
effectiveness is characterized by lowering the actuator gain with re- no-damage case, ξ(x) = 0.
spect to its nominal value. Different types of actuator failures are The objective of this paper is to design a controller that will
shown in Fig. 2. accommodate actuator failures and control effector damage even
Different types of actuator failures can be parameterized as while compensating for the effect of the disturbance. Hence the
follows: following control objective will be considered:


 u ci (t), ki (t) = 1, for all t ≥ 0, no-failure case



 ki (t)u ci (t), 0 < i ≤ ki (t) < 1, for all t ≥ tFi , loss of effectiveness
u i (t) = 0, ki (t) = 1, for all t ≥ tFi , float type of failure



 u ci (tFi ), ki (t) = 1, for all t ≥ tFi , lock-in-place failure

(u i )min or (u i )max , ki (t) = 1, for all t ≥ tFi , hard-over failure

where tFi denotes the time instant of failure of the ith actuator, ki Control objective 2: Design a control input u c (t) such that
denotes its effectiveness coefficient such that ki ∈ [ ki , 1], and ki > 0 x(t) − xm (t) ≤ for all time despite the effect of the disturbance
denotes its minimum effectiveness. ξ(x(t)), actuator failures, and control effector damage.
In the case of first-order actuator dynamics, the class of failures
just described can be modeled as follows: V. Simultaneous Failure and Damage Accommodation
and Disturbance Rejection
u̇ i = −σi λi (u i − ki u ci ) (22) To design a reconfigurable controller that effectively compensates
for the effect of LIP, LOE, HOF, or float failures, the expression (22)
where λi > 0; σi (t) = 1 in the case of no failure, and σi (t) = 0,
is first rewritten as
u(tFi ) = ū i when the failure occurs at t = tFi , where tFi denotes the
time of failure of the ith actuator. Hence in the case of failure at tFi , u̇ i = −λi (u i − σi ki u ci ) + λi (1 − σi )u i (25)
906 BOŠKOVIĆ, BERGSTROM, AND MEHRA

The preceding expression is next divided by λi , and singular pertur- ˆ = diag[σ̂1


 σ̂2 . . . σ̂m ] (36)
bation arguments are used to obtain
G̃ 0 (x) = [g1 (x)σ̂1 g2 (x)σ̂2 . . . gm (x)σ̂m ] (37)
ui ∼
= σi ki u ci + (1 − σi )ū i ,
where ū i = u i (tFi ), that is, ū i is the value at which the actuator has and let U = diag[u] and d̂ = diag[ D̂]. To ensure the invertibility of
locked at the time of failure. The preceding equation is next substi- G̃ 0 , at least p of the estimates σi are kept at value one for all time
tuted into Eq. (2) to obtain (compare assumption 1e).
Now the following theorem is considered.
ẋ1 = x2 (26) Theorem 1: The following control law for the system (28) and
m (29):
ẋ2 = f (x) + gi (x)di [σi ki u ci + (1 − σi )ū i ] + ξ(x) (27)  −1
i =1
u c = D̂ K̂ G̃ 0T G̃ 0 D̂ 2 K̂ 2 G̃ 0T [− f (x) + Am x + Bm r

It is seen that the presence of ū i in the preceding expression in- ˆ − ξ̂ ]


− G(x)(I − )u
troduces additional uncertainty into the model. For this reason the
where the estimates are adjusted by using
following assertion is considered.
Assertion 1: If σ ∈ {0, 1}, then (1 − σi )(ū i − u i ) ≡ 0. σ̂˙ i = φ̇σ i = Proj[0,1] {γσ i (u i − k̂i u ci )êi }, i = 1, 2, . . ., m (38)
Proof: The proof follows trivially because the assertion is true
Downloaded by INDIAN INSTITUTE OF TECHNOLOGY on April 3, 2019 | http://arc.aiaa.org | DOI: 10.2514/1.11272

for σi = 1, whereas for σi = 0, which happens at tFi , one has that


k̂˙ i = φ̇ki = Proj[ ki ,1] {−γki u ci êi }, i = 1, 2, . . ., m (39)
u i (tFi ) = ū i . 
Based on the result of the preceding assertion, the model (27) is
now rewritten as d̂i = Proj[ di ,1] {1 − sign[u i ê T P̄ gi (x)]} (40)
 

ẋ1 = x2 (28) m n
ξ̂i = − ci + di ϕi (x) + |gi (x)ωi | sign pji ê j

m i =1 j =1
ẋ2 = f (x) + gi (x)di [σi ki u ci + (1 − σi )u i ] + ξ(x) (29)
i =1 i = 1, 2, . . ., p (41)

Expression (29) describes the plant under a class of potential failures where γσ i > 0, γki > 0, and γdi > 0, ensures the boundedness of all
and damage described in the preceding section and will be used in signals in the system and, in addition, limt → ∞ [x(t) − xm (t)] = 0,
the FDIR design as described next. and limt → ∞ êui (t) = 0, i = 1, 2, . . ., m. (In the preceding equations
Proj{ · }denotes the projection operator14 the role of which is to
project the estimates σ̂i , k̂i , and d̂i to the intervals [0, 1], [ ki , 1],
A. Decentralized Adaptive FDI Observer
and [ di , 1], respectively.)
The decentralized FDI observer for estimating actuator failure- Before proceeding to the proof of the theorem, the following
related parameters is first built in the form assertion is considered.
Assertion 2: Let d ∈ [ d , 1] and
û˙ i = −σ̂i λ(u i − k̂i u ci ) − λ0 (û i − u i ) (30)
d̂ = Proj[ d ,1] {1 − sign(ω)} (42)
where λ0 > 0.
Let êui = û i − u i . Upon subtracting Eq. (22) from Eq. (30), one Then ω(d̂ − d) ≤ 0, for all d̂ ∈ [ d , 1], all d ∈ [ d , 1], and all ω.
obtains Proof: The proof is based on considering two cases. The first
case is when 1 − sign(ω) < 1, which implies that d̂ = 1 − sign(ω).
ê˙ ui = −λ0 êui − λφσ i u i + λ(σ̂i k̂i u ci − σi ki u ci )
In such a case one has that
= −λ0 êui + λσi φki u ci − λ(u i − k̂i u ci )φσ i d̂ω = ω − |ω|, ωd̂ − ωd = ω(1 − d) − |ω| ≤ 0
= −λ0 êui + φiT ωi (31) In the second case one has that 1 − sign(ω) ≥ 1, which also implies
that d̂ = 1. From this inequality it also follows that sign(ω) ≤ 0 and
where φσ i = σ̂i − σi , φki = k̂i − ki , φi = [φσi φki ]T , and ωi = ω ≤ 0; hence,
λ[−(u i − k̂i u ci ) σi u ci ]T .
−|ω| + |ω|d ≤ 0
B. Global Robust FDI Observer because d ≤ 1, which completes the proof. 
The next step is to design an observer, based on expressions (28) Proof of Theorem 1: The proof is divided into two parts. In the
and (29), for estimation of the effect of damage and state-dependent first part the objective is to demonstrate that the decentralized es-
disturbances: timation of actuator failure-related parameters yields bounded esti-
mation errors, whereas in the second part it will be shown that the
x̂˙ 1 = x̂2 (32) preceding algorithms ensure effective compensation for the effect
of state-dependent disturbances, control effector damage, and the
x̂˙ 2 = Am (x̂ − x) + f (x) coupling between the decentralized estimation subsystem and the

m  damage and disturbance estimation subsystem.
+ gi (x)d̂i [σ̂i k̂i u ci + (1 − σ̂i )u i ] + ξ̂ (33) The following tentative Lyapunov function is chosen first:
i =1  
V (êui , φσ i , φki ) = 1
2
êui2 + λ[φσ i /γσ i + σi (φki /γki )]
where the estimates σ̂i and k̂i are generated by the decentralized FDI
subsystem. Hence, V is positive semidefinite if σi = 0. Its first derivative along
the motions of the system yields
Let φdi = d̂i − di , ê = x̂ − x, and
V̇ (êui , φσ i , φki ) ≤ −λ0 êui2 ≤ 0
D̂ = diag[d̂1 d̂2 . . . d̂m ] (34)
The latter inequality holds regardless of the value of σi . Because
K̂ = diag[k̂1 k̂2 . . . m̂ m ] (35) adaptive algorithms with projection are used, k̂i will always be
BOŠKOVIĆ, BERGSTROM, AND MEHRA 907

bounded, along with σ̂i . From the derivative of V , one can con- Using expression (41), the derivative of V is now
clude that êui is bounded. It can also be readily demonstrated that  
êui ∈ L2 . However, it cannot yet be concluded that this error tends 1 p m n

to zero asymptotically. V̇ (ê) ≤ − ê T Q ê + ci + di ϕi (x) + |gi (x)ωi | pji ê j


2 i =1 i =1 j =1
The global FDI observer (32) and (33) is now written as
p  
x̂˙ 1 = x̂2 (43)
m
n
− ci + di ϕi (x) +
|gi (x)ωi | pji ê j
x̂˙ 2 = Am (x̂ − x) + f (x) + G(x) D̂[
ˆ K̂ u c + (I − 
ˆ i )u] + ξ̂ i =1 i =1 j =1

(44) 1
≤ − ê T Q ê < 0
2
Because D̂, K̂ , and  ˆ are diagonal, one has that G(x) D̂ 
ˆ K̂ =
G(x) ˆ D̂ K̂ = G̃ 0 (x) D̂ K̂ . hence, ê is bounded, and limt → ∞ ê(t) = 0.
Substituting the control law (38) into the preceding observer equa- Because it has been already shown that limt → ∞ [x̂(t) − xm (t)] = 0
tion results in the following closed-loop observer system: and because em = ê + e, where e = x − xm denotes the tracking error,
it follows that x is bounded and that limt → ∞ e(t) = 0. It can now
x̂˙ 1 = x̂2 , x̂˙ 2 = Am x̂ + Bm r be readily verified that ê˙ ui is bounded, which, using the fact that
êui ∈ L∞ ∩ L2 , implies that limt → ∞ êui (t) = 0, i = 1, 2, . . . , m. 
Downloaded by INDIAN INSTITUTE OF TECHNOLOGY on April 3, 2019 | http://arc.aiaa.org | DOI: 10.2514/1.11272

Upon subtracting the reference model equations, it can be readily One can again use the approximation
shown that the error between the observer and the reference model
will tend to zero. Because the state of the reference model xm is
bounded, it follows that x̂ is bounded as well. However, it remains sign(η) ∼
= η/(|η| + δ)
to be shown that x is also bounded.
Recalling that ê = x̂ − x and êi = x̂i − xi , i = 1, 2, subtracting the where 0 < δ 1, to prevent chattering and ensure the UUB of the
expressions (28) and (29) from Eqs. (32) and (33) yields signals in the system.
ê˙ 1 = ê2 (45)

m VI. Simulations
ê˙ 2 = Am e + G(x)U (d̂ − d) + ξ̂ − ξ + gi (x)d̂i φiT ωi (46) In this section the results of performance evaluation of the robust
i =1 integrated FDIR scheme on a simulation of the Boeing’s TAFA are
presented. The results presented are for the case of lateral doublet.
where d = diag[D] and d̂ and U were defined earlier. Recent performance evaluations of a simplified version of the pro-
The expressions (45) and (46) can be rewritten in a compact form posed integrated FDIR system through F/A-18 piloted simulations
as are also briefly described.

m 
ê˙ = A0 ê + C0 GU (d̂ − d) + ξ̂ − ξ + gi (x)φi ωi (47) A. TAFA Simulations
i =1 The TAFA is a conceptual design of an advanced fighter configu-
where A0 and C0 were defined earlier. ration that blends an extensive suite of conventional and innovative
Let a tentative Lyapunov function be of the form control effectors to achieve high agility in a low observable design.
The TAFA is a single-engine, single-seat fighter designed for air-to-
V (ê) = 12 ê T P ê air and/or air-to-ground missions. In this paper the focus is on the
flight condition with altitude at the sea level, Mach number 0.9, and
where P = P T > 0 is a solution to the Lyapunov matrix equation an angle of attack of 1.65 deg.
A0T P + P A0 = −Q, where Q = Q T > 0. A linearized model of TAFA dynamics is of the form
The first derivative of V along the motions of Eq. (47) yields
1 ẋ = Ax + Bu (48)
V̇ (ê) = − ê T Q ê + ê T P̄ G(x)U (d̂ − d)
2

m  The states of the linearized TAFA dynamics are x =
+ ê P̄ ξ̂ − ξ +
T
gi (x)φi ωi [V q θ α h β p r φ ψ]T , where V de-
i =1 notes perturbed forward velocity, q denotes the perturbed pitch
rate, θ denotes the perturbed pitch angle, α denotes the perturbed
where P̄ = PC0 . Using the assertion 2 and the adjustment law for angle of attack, h denotes the perturbed altitude, β denote the
d̂ given by Eq. (40), the derivative of V becomes perturbed sideslip angle, p and r denote respectively the per-
 m  turbed roll and yaw rates, and φ and ψ denote respectively the
1 perturbed bank and yaw angles, where all perturbations are with
V̇ (ê) ≤ − ê T Q ê + ê T P̄ ξ̂ − ξ + gi (x)φi ωi
2 respect to a trim. The control inputs of TAFA are given as follows:
i =1
u 1 , perturbed left trailing-edge flap deflection (TEFL), rad; u 2 ,
The preceding expression can be rewritten as perturbed left canard deflection (CNDL), rad; u 3 , perturbed pitch
p 
m n  thrust vectoring nozzle deflection (NOZp), rad; u 4 , perturbed left
1 aft-body flap deflection (ABFL), rad; u 5 , perturbed left aileron de-
V̇ (ê) = − ê T Q ê + ξi − ξ̂i + gi (x)φi ωi pji ê j
2 flection (AILL), rad; u 6 , perturbed right aileron deflection (AILR),
i =1 i =1 j =1
rad; u 7 , perturbed right aft-body flap deflection (ABFR), rad; u 8 ,
p  m n  perturbed right trailing-edge flap deflection (TEFR), rad; u 9 , per-
1 turbed right canard deflection (CNDR), rad; u 10 , perturbed yaw
≤ − ê T Q ê + ci + di ϕi (x) + |gi (x)ωi | pji ê j
2 i =1 i =1 j =1
thrust-vectoring nozzle deflection (NOZy), rad; u 11 , perturbed dif-
ferential leading-edge flap deflection (DfLEF), rad; and u 12 , per-

p
n turbed continuous moldline deflection (DCMUP), rad.
+ ξ̂i pji ê j The linearized stability and control derivative matrices for TAFA
i =1 j =1 dynamics around this flight condition are of the form
908 BOŠKOVIĆ, BERGSTROM, AND MEHRA

 
−0.06 −27.25 −32.21 75.93 0 0 0 0 0 0
 0 −4.28 0 8.74 0 0 0 0 0 0
 
 0 1 0 0 0 0 0 0 0 0
 
 0 −1.86 0 0
 1 0 0 0 0 0 
 0 1004.8 −1004.8 0 0
 0 0 0 0 0 
A= 
 0 0 0 0 0 −0.12 0.03 −1 0.03 0
 
 0 0 0 0 0 −30.85 −5.81 0.03 0 0
 
 0 0 0 0 0 −11.67 0.04 −0.55 0 0
 
 0 0 0 0 0 0 1 0.03 0 0
0 0 0 0 0 0 0 1 0 0
 
−0.31 20.38 0 −5.18 0.12 0.12 −5.18 −0.31 20.38 0 0 0
 −14.07 14.93 −3.94 0 −6.23 −6.23 0 −14.07 14.93 0 0 0 
 
 0 0 0 0 0 0 0 0 0 0 0 0 
 
 −0.15 −0.03 −0.02 
Downloaded by INDIAN INSTITUTE OF TECHNOLOGY on April 3, 2019 | http://arc.aiaa.org | DOI: 10.2514/1.11272

 0 −0.06 −0.06 0 −0.15 −0.03 0 0 0 


 0 
 0 0 0 0 0 0 0 0 0 0 0 
B= 
 −0.02 0.01 0 0.02 0 0 −0.02 0.02 −0.01 0.02 −0.01 0 
 
 49.9 3.88 0 0.15 34.72 −34.72 −0.15 −49.9 −3.88 0.12 104.73 −1034.6
 
 2.45 3.6 0 −4.24 0.46 −0.46 4.24 −2.45 −3.6 −3.44 4.74 2.9 · 104 
 
 0 0 0 0 0 0 0 0 0 0 0 0 
0 0 0 0 0 0 0 0 0 0 0 0

Table 1 Position and rate limits on the control effectors 4. Simulation Scenario
Position limit, Rate limit, The simulation scenario consists of a lateral doublet (maximum
Effector deg deg/s magnitude φ = 60 deg). No command in ψ is specified, which makes
Left trailing-edge flap [−30, 45] [−90, 90] this a demanding maneuver for the TAFA. At time t = 3.5 s, the air-
Left canard [−80, 10] [−70, 70] craft undergoes multiple simultaneous failures and damage. The
NOZp - pitch thrust-vectoring nozzle [−30, 30] [−70,70] complete failure scenario consists of 1) two lock-in-place failures
Left aftbody flap [0, 90] [−120, 120] of the left canard (CNDL) and left aftbody flap (ABFL); 2) sur-
Left aileron [−30, 30] [−120, 120] face damage to the left trailing-edge flap (TEFL) and left aileron
Right aileron [−30, 30] [−120, 120] (AILL) such that dTEFL = 0.2 and dAILL = 0.2; and 3) a disturbance
Right aftbody flap [0, 90] [−120, 120] input caused by the damage. The damage-generated disturbance is
Right trailing-edge flap [−30, 45] [−90,90] assumed to cause nonlinear effects in the angular rates p, q, and r ,
Right canard [−80, 10] [−70, 70]
of the form
NOZy-yaw thrust-vectoring nozzle [−30, 30] [−60, 60]
Differential leading-edge flaps [−30, 30] [−70, 70]
ξq = −4α 2 − 10β 2 + −2 p 2 − 4q 2 − 8r 2 − 2 pq − 2 pr − 4qr

1. Actuator Dynamics, Position, and Rate Limits (49)


All actuators are characterized by first-order dynamics of the form
ξ p = −10α − 6β − 2 p − 6q − 16r − 4 pq − 4 pr − 2qr
2 2 2 2 2

u = [40/(s + 40)]u c
(50)
where u is the actual position of the effector and u c is the position
command for that effector generated by the controller. ξr = −2α 2 − 10β 2 − 6 p 2 − 10q 2 − 4r 2 − 8 pq − 8 pr − 2qr
The input vector u = [u 1 . . . u 12 ] defined earlier is expressed in
terms deflections of the left and right flaps. Their position and rate (51)
limits are listed in Table 1.
It is seen that the damage is assumed to result in a large nonlinearity.
2. Medium-Fidelity Nonlinear Simulation of TAFA Dynamics
This type of disturbance was chosen because the p, q, r squared
The simulation takes into account the rotation from the attitude and cross terms represent the type of effect that arises when damage
angles to the body angular rates, Coriolis acceleration, and the ef- causes asymmetry in the vehicle kinematics. The α and β terms
fects of gravity, whereas aerodynamic forces and moments are ap- were included to model aerodynamic effects caused by the damage.
proximated by the terms from the stability and control derivative In the simulations, because the main maneuver is in p, the p 2 term
matrices. All position and rate limits and actuator dynamics are in- generally dominates.
cluded in the simulation, along with actuator failures (LIP, LOE)
and control effector damage.
5. Controller Parameters
3. FDIR System Design and Implementation The bounds on the nonlinearity are assumed to be of the form
The nonlinear model used for simulation of TAFA dynamics was
transformed along the lines described in Sec. II, and, when the ac- ϕ p = ϕq = ϕr = 20(|α| + |β| + | p| + |q| + |r |)2 (52)
tuator failures and control effector damage are included, has the
form (23) and (24). The transformed system states are θ, φ, ψ, q, p, which will always be larger in magnitude than the true disturbance,
r , V , α, and β. The controller is given by Eq. (38), and parameter while having a similar form. The adaptive gains were chosen as
adjustment laws are given by Eqs. (38–41). γσ i = 500 and γki = 5. The adaptive observer gain is λ = 40.
BOŠKOVIĆ, BERGSTROM, AND MEHRA 909

6. Reference Model with resulting disturbance, as just described. The baseline controller
The reference model matrices are given next. It is noted that, for that has no reconfiguration capability provides poor performance in
the desired dynamics of V , α, and β, the elements of the Am and case 2, and is not capable of stabilizing the system in cases 3 and
Bm matrices are chosen from the corresponding diagonal elements 4. The simulations with the proposed integrated FDIR system are
of the A matrix: detailed next.
  Nominal case. First a nominal simulation is presented where
A11 0 0 0 0 0 0 0 0 0
there are no failures or damage. The proposed integrated FDIR sys-
 0 −7 −25 0 0 0 0 0 0 0
  tem is running throughout the simulation and does not generate any
 0 1 0 0 0 0 0 0 0 0 false alarms. (The performance is identical to that of an unrecon-
 
 0 0 figurable baseline controller.) The resulting response is shown in
 1 0 A44 0 0 0 0 0 
 0  Figs. 3–5.
 0 A53 A54 0 0 0 0 0 0 Actuator lock-in-place failures only. In this case the aircraft
Am =  
 0 0 0 0 0 A66 0 −1 0 0 suffers lock-in-place failures of the left canard (CNDL) and left
  aftbody flap (ABFL) at t = 3.5 s. The resulting response is shown
 0 0 0 0 0 0 −7 0 −25 0
  in Fig. 6–9. It is seen that the system is stabilized, and the resulting
 0 0 0 0 0 0 0 −7 0 −25
  response is comparable to that obtained in the no-failure case.
 0 0 0 0 0 0 1 0 0 0 Control effector damage and disturbance only. In this case
0 0 0 0 0 0 0 1 0 0 the left trailing-edge flap (TEFL) and left aileron (AILL) undergo
Downloaded by INDIAN INSTITUTE OF TECHNOLOGY on April 3, 2019 | http://arc.aiaa.org | DOI: 10.2514/1.11272

damage at t = 3.5 s, and their remaining effectiveness is 0.2, that


  is, dTEFL = 0.2 and dAILL = 0.2. The damage-generated disturbance
−A11 0 0 0 0 0 is of the form (49–51). The response of the system is shown in
 0 0 25 0 0 0 Figs. 10–12. It is seen that, after a transient, the system stabilizes at
 
 0 0 0 0 0 0 trim, and that the tracking in φ is accurate.
 
 0 −A44 0
Simultaneous actuator lock in place and control effector damage.
 0 0 0  This is the most challenging scenario, in which the two preceding
 0 0
 0 0 0 0  failure scenarios are combined. In this case, the aircraft suffers two
Bm =  
 0 0 0 −A66 0 0 lock-in-place actuator failures, two cases of control effector damage,
  and a large resulting disturbance. The response of the system is
 0 0 0 0 25 0
  shown in Figs. 13–16. It is seen that the response is good and that
 0 0 0 0 0 25
  the actuator lock in place is accurately identified.
 0 0 0 0 0 0
B. F/A-18 Simulations
0 0 0 0 0 0
An integral part of the proposed robust integrated FDIR sys-
tem is the decentralized FDIR algorithms, also referred to as the
7. Simulations FLARE (Fast on-Line Actuator Recovery Enhancement) system,
All simulations are run with the proposed integrated FDIR sys- developed under a NASA Dryden Small Business Innovation Re-
tem. The following cases are included: 1) nominal (no-failure/no- search (SBIR) project15 for the case of flight-critical actuator fail-
damage) case; 2) actuator lock-in-place only (no damage/no distur- ures. The FLARE algorithms were recently evaluated at Boeing
bances); 3) control effector damage with a resulting disturbance through high-fidelity and piloted simulations of the F/A-18C/D air-
(no actuator failures); and 4) the full-failure scenario, in which craft under lock-in-place and hard-over actuator failures.16 Because
there are simultaneous actuator failures and control effector damage the FLARE system was developed for a linearized model of F/A-18

Fig. 3 Longitudinal state response with the proposed FDIR system in the no-failure/no-damage case.
910 BOŠKOVIĆ, BERGSTROM, AND MEHRA
Downloaded by INDIAN INSTITUTE OF TECHNOLOGY on April 3, 2019 | http://arc.aiaa.org | DOI: 10.2514/1.11272

Fig. 4 Lateral state response with the proposed FDIR system in the no-failure/no-damage case.

Fig. 5 Control input response with the proposed FDIR system in the no-failure/no-damage case.
BOŠKOVIĆ, BERGSTROM, AND MEHRA 911
Downloaded by INDIAN INSTITUTE OF TECHNOLOGY on April 3, 2019 | http://arc.aiaa.org | DOI: 10.2514/1.11272

Fig. 6 Longitudinal state response with the proposed FDIR system in the case of actuator lock-in-place.

Fig. 7 Lateral state response with the proposed FDIR system in the case of actuator lock-in-place.

dynamics, the scheme also includes variable-structure compensa- rating scale. The results of the piloted simulations in the case of
tion of unmodeled dynamics caused by linearization. The design aileron, rudder, and stabilator failures are summarized in Fig. 17.
was initially carried out in MATLAB® by using a low-fidelity F/A- The aileron and rudder underwent lock-in-place failures at 0 and
18 simulation and was subsequently evaluated on a high-fidelity 15 deg and hard-over failures to their position limits of ±30 deg,
F/A-18 simulation at Boeing. Following successful evaluations on while the stabilator was locked at 0, 3, and 6 deg. It is seen
the high-fidelity simulator, the FLARE algorithms were imported that the FLARE system substantially improves the flight perfor-
into the piloted flight simulator at Boeing, and their performance mance under flight-critical failures, and, in some cases, the failure
was evaluated by a pilot during longitudinal and lateral doublets, a and its subsequent accommodation were barely noticeable by the
4-g turn, a 360-deg roll, and tracking tasks using Cooper–Harper pilot.
912 BOŠKOVIĆ, BERGSTROM, AND MEHRA
Downloaded by INDIAN INSTITUTE OF TECHNOLOGY on April 3, 2019 | http://arc.aiaa.org | DOI: 10.2514/1.11272

Fig. 8 Control input response with the proposed FDIR system in the case of actuator lock-in-place.

Fig. 9 Actuator failure parameter estimates in the case of actuator lock-in-place.


BOŠKOVIĆ, BERGSTROM, AND MEHRA 913
Downloaded by INDIAN INSTITUTE OF TECHNOLOGY on April 3, 2019 | http://arc.aiaa.org | DOI: 10.2514/1.11272

Fig. 10 Longitudinal state response with the proposed FDIR system in the case of control effector damage.

Fig. 11 Lateral state response with the proposed FDIR system in the case of control effector damage.
914 BOŠKOVIĆ, BERGSTROM, AND MEHRA
Downloaded by INDIAN INSTITUTE OF TECHNOLOGY on April 3, 2019 | http://arc.aiaa.org | DOI: 10.2514/1.11272

Fig. 12 Control input response with the proposed FDIR system in the case of control effector damage.

Fig. 13 Longitudinal state response with the proposed FDIR system in the case of simultaneous actuator failures and control effector damage.
BOŠKOVIĆ, BERGSTROM, AND MEHRA 915
Downloaded by INDIAN INSTITUTE OF TECHNOLOGY on April 3, 2019 | http://arc.aiaa.org | DOI: 10.2514/1.11272

Fig. 14 Lateral state response with the proposed FDIR system in the case of simultaneous actuator failures and control effector damage.

Fig. 15 Control input response with the proposed FDIR system in the case of simultaneous actuator failures and control effector damage.
916 BOŠKOVIĆ, BERGSTROM, AND MEHRA
Downloaded by INDIAN INSTITUTE OF TECHNOLOGY on April 3, 2019 | http://arc.aiaa.org | DOI: 10.2514/1.11272

Fig. 16 Actuator failure parameter estimates with the proposed FDIR system in the case of simultaneous actuator failures and control effector
damage.

Fig. 17 Results of Cooper–Harper evaluation of FLARE under aileron, rudder, and stabilator failures.

VII. Conclusions Acknowledgment


In the paper, a robust integrated fault-tolerant flight control sys- This research was partially supported by the NASA Dryden Flight
tem that accommodates for different types of actuator failures and Research Center under Contract NAS4-02017 to Scientific Systems
control effector damage is presented even while rejecting state- Company.
dependent disturbances. It is shown that variable structure laws for
the adjustment of disturbance estimates yield a stable system despite
simultaneous presence of failures, damage, and disturbances. The
References
1 Bodson, M., and Groszkiewicz, J., “Multivariable Adaptive Algorithms
properties of the proposed algorithms are illustrated on a medium-
fidelity simulation of Boeing’s Tailless Advanced Fighter Aircraft. for Reconfigurable Flight Control,” IEEE Transactions on Control Systems
The implementation of the FLARE system, which represents an Technology, Vol. 5, No. 2, 1997, pp. 217–229.
2 Ahmed-Zaid, F., Ioannou, P., Gousman, K., and Rooney, R., “Accom-
integral part of the proposed failure, detection, identification, and
reconfiguration system, on a piloted simulator of F/A-18 aircraft is modation of Failures in the F-16 Aircraft Using Adaptive Control,” IEEE
Control Systems Magazine, Vol. 11, No. 1, 1991, pp. 73–78.
also described. 3 Monaco, J. D., Ward, D., Barron, R., and Bird, R., “Implementation
The results presented in the paper appear fairly promising con- and Flight Test Assessment of an Adaptive, Reconfigurable Flight Control
sidering that the proposed system achieves acceptable performance System,” AIAA Paper 97-3738, Aug. 1997.
despite two locked actuators, two damaged control effectors, and a 4 Chandler, P., Pachter, M., and Mears, M., “System Identification for
large disturbance affecting all three angular rates. Future plans in- Adaptive and Reconfigurable Control,” Journal of Guidance, Control, and
clude extensive testing of the proposed algorithms on a high-fidelity Dynamics, Vol. 18, No. 3, 1995, pp. 516–524.
simulator and their implementation to other manned and unmanned 5 Calise, A., Lee, S., and Sharma, M., “Direct Adaptive Reconfigurable
vehicles. Control of a Tailless Fighter Aircraft,” Proceedings of the 1998 AIAA
BOŠKOVIĆ, BERGSTROM, AND MEHRA 917

Guidance, Navigation, and Control Conference, Vol. 1, Reston, VA, 1998, 1998 AIAA Guidance, Navigation, and Control Conference, Vol. 1, Reston,
pp. 88–97. VA, 1998, pp. 118–126.
6 Brinker, J., and Wise, K., “Reconfigurable Flight Control of a Tailless 12 Bošković, J. D., and Mehra, R. K., “Robust Fault-Tolerant Control De-
Advanced Fighter Aircraft,” Proceedings of the 1998 AIAA Guidance, Navi- sign for Aircraft Under State-Dependent Disturbances,” AIAA Paper 2003-
gation, and Control Conference, Vol. 1, AIAA, Reston, VA, 1998, pp. 75–87. 5490, Aug. 2003.
7 Brinker, J., and Wise, K., “Flight Testing of a Reconfigurable Flight 13 Hsu, L., and Costa, R., “Variable Structure Model Reference Adaptive
Control Law on the X-36 Tailless Fighter Aircraft,” AIAA Paper 2000-3941, Control Using Only I/O Measurements,” International Journal of Control,
Aug. 2000. Vol. 49, No. 2, 1989, pp. 399–416.
8 Bošković, J. D., and Mehra, R. K., “A Decentralized Scheme for Au- 14 Bošković, J. D., and Mehra, R. K., “A Multiple Model Adaptive Flight
tonomous Compensation of Multiple Simultaneous Flight-Critical Failures,” Control Scheme for Accommodation of Actuator Failures,” Journal of Guid-
AIAA Paper 2002-4453, Aug. 2002. ance, Control, and Dynamics, Vol. 25, No. 4, 2002, pp. 712–724.
9 Bošković, J. D., and Mehra, R. K., “An Adaptive Scheme for Compensa- 15 Bošković, J. D., Bergstrom, S. E., and Mehra, R. K., “Aircraft Prognos-
tion of Loss of Effectiveness of Flight Control Effectors,” Proceedings of the tics and Health Management (PHM) and Adaptive Reconfigurable Control
40th IEEE Conference on Decision and Control, Vol. 3, Inst. of Electrical (ARC) System,” NASA DFRC Phase II SBIR, Final Rept., Contract NAS4-
and Electronics Engineers, Piscataway, NJ, 2001, pp. 2448–2453. 02017, March 2004.
10 Bošković, J. D., and Mehra, R. K., “Intelligent Adaptive Control of a 16 Bošković, J. D., Bergstrom, S. E., Mehra, R. K., Urnes, J. M.,
Tailless Advanced Fighter Aircraft Under Wing Damage,” Journal of Guid- Sr., Hood, M., and Lin, Y., “Performance Evaluation of an Integrated Retrofit
ance, Control, and Dynamics, Vol. 23, No. 5, 2000, pp. 876–884. Failure Detection, Identification and Reconfiguration (FDIR) System Using
11 Wise, K., and Sedwick, J., “Stability Analysis of Reconfigurable and High-Fidelity and Piloted Simulations,” Society of Automotive Engineers,
Gain Scheduled Flight Control System Using LMIs,” Proceedings of the SAE Paper 2004-01-3115, Nov. 2004.
Downloaded by INDIAN INSTITUTE OF TECHNOLOGY on April 3, 2019 | http://arc.aiaa.org | DOI: 10.2514/1.11272
This article has been cited by:

1. Liang Zhang, Changzhu Wei, Rong Wu, Naigang Cui. 2019. Adaptive fault-tolerant control for a VTVL reusable launch
vehicle. Acta Astronautica . [Crossref]
2. Bingqian Li, Wenhan Dong, Chao Xiong. 2019. Robust Actuator‐Fault‐Tolerant Control System Based on Sliding‐Mode
Observer for Thrust‐Vectoring Aircrafts. Asian Journal of Control 21:1, 236-247. [Crossref]
3. Zhongzheng Zhang, Dong Ye, Bing Xiao, Zhaowei Sun. 2019. Third‐order sliding mode fault‐tolerant control for satellites
based on iterative learning observer. Asian Journal of Control 21:1, 43-51. [Crossref]
4. Hupo Ouyang, Yan Lin. 2018. Supervisory adaptive fault-tolerant control against actuator failures with application to an
aircraft. International Journal of Robust and Nonlinear Control 28:2, 536-551. [Crossref]
5. Jovan Boskovic, Joseph A. Jackson, Richard Wise, Nhan T. Nguyen. Adaptive Output Feedback Control of Elastically
Shaped Aircraft . [Citation] [PDF] [PDF Plus]
6. Jovan D. Bošković, Joseph A. Jackson. 2017. Adaptive Flight Control Under Actuation and Sensor Failures and Slow-
Downloaded by INDIAN INSTITUTE OF TECHNOLOGY on April 3, 2019 | http://arc.aiaa.org | DOI: 10.2514/1.11272

Engine Dynamics. Journal of Guidance, Control, and Dynamics 40:4, 905-919. [Abstract] [Full Text] [PDF] [PDF Plus]
7. Jovan Boskovic, Joseph A. Jackson, Richard Wise. Implementation & Flight Testing of the IMPACT System for
Autonomous ISR . [Citation] [PDF] [PDF Plus]
8. Jovan Boskovic, Richard Wise, Joseph A. Jackson, Nhan T. Nguyen. A Flutter Suppression and Drag Optimization
Approach for Flexible Aircraft . [Citation] [PDF] [PDF Plus]
9. Bin Zhang, Tao Li, Gongfei. Multi-bound-dependent fault-tolerant control for flexible spacecraft under partial loss of
actuator effectiveness 116-122. [Crossref]
10. Ehsan Mohammadi, Morteza Montazeri-Gh. 2016. Active Fault Tolerant Control with self-enrichment capability for gas
turbine engines. Aerospace Science and Technology 56, 70-89. [Crossref]
11. Liang Zhang, Changzhu Wei, Liang Jing, Naigang Cui. Heavy lift launch vehicle technology of adaptive augmented fault
tolerant control 1587-1593. [Crossref]
12. Bing Xiao, Shen Yin. 2016. Velocity-Free Fault-Tolerant and Uncertainty Attenuation Control for a Class of Nonlinear
Systems. IEEE Transactions on Industrial Electronics 63:7, 4400-4411. [Crossref]
13. Shen Yin, Bing Xiao, Steven X. Ding, Donghua Zhou. 2016. A Review on Recent Development of Spacecraft Attitude
Fault Tolerant Control System. IEEE Transactions on Industrial Electronics 63:5, 3311-3320. [Crossref]
14. Jovan Boskovic, Richard Wise, Joseph A. Jackson. Drag Identification and Reduction Technology (DIRECT) for
Commercial Aircraft . [Citation] [PDF] [PDF Plus]
15. Jovan Boskovic, Joseph A. Jackson. An Innovative Approach to Air Data Sensor FDIR for Commercial Aircraft . [Citation]
[PDF] [PDF Plus]
16. Dezhi Xu, Bin Jiang, Peng Shi. 2015. Robust NSV Fault-Tolerant Control System Design Against Actuator Faults
and Control Surface Damage Under Actuator Dynamics. IEEE Transactions on Industrial Electronics 62:9, 5919-5928.
[Crossref]
17. Jianping Cai, Changyun Wen, Hongye Su. 2015. Adaptive inverse control for parametric strict feedback systems with
unknown failures of hysteretic actuators. International Journal of Robust and Nonlinear Control 25:6, 824-841. [Crossref]
18. Tawfiqur Rahman, Hao Zhou, Liang Yang, Wanchun Chen. 2015. Pseudospectral Model Predictive Control for Exo-
atmospheric Guidance. International Journal of Aeronautical and Space Sciences 16:1, 64-76. [Crossref]
19. Bing Xiao, Qinglei Hu, Danwei Wang. 2015. Spacecraft Attitude Fault Tolerant Control with Terminal Sliding-Mode
Observer. Journal of Aerospace Engineering 28:1, 04014055. [Crossref]
20. Jovan D. Bošković. 2014. A new decentralized retrofit adaptive fault-tolerant flight control design. International Journal of
Adaptive Control and Signal Processing 28:9, 778-797. [Crossref]
21. . Robust Fault Estimation 219-256. [Crossref]
22. . Robust Fault Detection 49-134. [Crossref]
23. Bing Xiao, Qinglei Hu, Michael I. Friswell. 2013. Active fault-tolerant attitude control for flexible spacecraft with loss of
actuator effectiveness. International Journal of Adaptive Control and Signal Processing 27:11, 925-943. [Crossref]
24. Jovan Boskovic, Joseph A. Jackson, Raman Mehra. Robust Adaptive Control in the Presence of Unmodeled Actuator
Dynamics . [Citation] [PDF] [PDF Plus]
25. Dezhi Xu, Bin Jiang, Hongtao Liu, Peng Shi. 2013. Decentralized asymptotic fault tolerant control of near space vehicle
with high order actuator dynamics. Journal of the Franklin Institute . [Crossref]
26. Chunyan Gao, Guangren Duan. 2012. Robust Adaptive Fault Estimation for a Class of Nonlinear Systems Subject to
Multiplicative Faults. Circuits, Systems, and Signal Processing 31:6, 2035-2046. [Crossref]
27. Dezhi Xu, Bin Jiang, Hongtao Liu. Fast Robust Asymptotic Fault Accommodation for Reentry Near Space Vehicles .
[Citation] [PDF] [PDF Plus]
28. Yuying Guo, Youmin Zhang, Bin Jiang, Zhengwei Zhu. Multiple-model-based adaptive reconfiguration control of state
delayed systems with actuator faults 397-402. [Crossref]
29. Sung‐Sik Shin, Byoung‐Mun Min, Min‐Jea Tahk. 2012. Retrofit Flight Control Using an Adaptive Chebyshev Function
Approximator. Journal of Aerospace Engineering 147. [Crossref]
30. Jie Chen, Hongchao Zhao, Yong Liang. Unknown control direction hypersonic aircraft adaptive fault-tolerant control
309-314. [Crossref]
31. Y Guo, B Jiang, Y Zhang, Z-W Zhu. 2012. Multiple model adaptive reconfiguration control of state delayed systems.
Downloaded by INDIAN INSTITUTE OF TECHNOLOGY on April 3, 2019 | http://arc.aiaa.org | DOI: 10.2514/1.11272

Proceedings of the Institution of Mechanical Engineers, Part I: Journal of Systems and Control Engineering 226:3, 325-337.
[Crossref]
32. Qinglei Hu. 2012. Adaptive Nonlinear Proportional-Derivative Type Fault Tolerant Control for Flexible Spacecraft
Attitude Maneuvers under Bounded Disturbances. Journal of Aerospace Engineering 25:2, 178. [Crossref]
33. Jianping Cai, Changyun Wen, Hongye Su, Zhitao Liu, Xiangbin Liu. Robust adaptive failure compensation of hysteretic
actuators for parametric strict feedback systems 7988-7993. [Crossref]
34. Yigang Sun, Daming Ren. Research on Aero-Engine Sensor Failure Diagnosis with Dual Redundant Predictors Based on
Neutral Network 1-4. [Crossref]
35. Jungho Kim, Tailian Chen. Heat Transfer Enhancement: Phase Change, Geometry, and Jets/Sprays . [Crossref]
36. Youdan Kim, Sungwan Kim. Reconfigurable Flight Control . [Crossref]
37. Nicolas Léchevin, Camille A. Rabbath, Chun-Yi Su. Integrated Health Monitoring for Multiple Air Vehicles . [Crossref]
38. Israel Lopez, Nesrin Sarigul-Klijn. 2010. A review of uncertainty in flight vehicle structural damage monitoring, diagnosis
and control: Challenges and opportunities. Progress in Aerospace Sciences 46:7, 247-273. [Crossref]
39. Jovan D. Boskovic, Raman K. Mehra. 2010. A Decentralized Fault-Tolerant Control System for Accommodation of Failures
in Higher-Order Flight Control Actuators. IEEE Transactions on Control Systems Technology 18:5, 1103-1115. [Crossref]
40. Jovan Boskovic, Nathan Knoebel, Joseph Jackson. An Initial Study of Pilot-Adaptive Controller Interactions in Flight
Control . [Citation] [PDF] [PDF Plus]
41. In-Seok Yang, Young-Jin Kim, Dong-Ik Lee. 2010. Actuator Failure Diagnosis and Accommodation Using Sliding Mode
Control for Submersible Vehicle. Journal of Institute of Control, Robotics and Systems 16:7, 661-667. [Crossref]
42. Jovan D. Boskovic, Joseph A. Jackson, Raman K. Mehra, Nhan T. Nguyen. 2009. Multiple-Model Adaptive Fault-Tolerant
Control of a Planetary Lander. Journal of Guidance, Control, and Dynamics 32:6, 1812-1826. [Citation] [PDF] [PDF Plus]
43. Jovan Boskovic, Nathan Knoebel. Adaptive Control of Plants with Actuators Operating on Different Time Scales .
[Citation] [PDF] [PDF Plus]
44. Jovan Boskovic, Joshua Redding, Nathan Knoebel. An Adaptive Fault Management (AFM) System for Resilient Flight
Control . [Citation] [PDF] [PDF Plus]
45. Jovan Boskovic, Joshua Redding. An Autonomous Carrier Landing System for Unmannned Aerial Vehicles . [Citation]
[PDF] [PDF Plus]
46. Jovan Boskovic, Joshua Redding. Accommodation of Control Actuator Failures in Morphing Aircraft . [Citation] [PDF]
[PDF Plus]
47. N. Lechevin, C.A. Rabbath. 2009. Decentralized Detection of a Class of Non-Abrupt Faults With Application to
Formations of Unmanned Airships. IEEE Transactions on Control Systems Technology 17:2, 484-493. [Crossref]
48. Yuta KOBAYASHI, Masaki TAKAHASHI. 2009. Design of Intelligent Fault-Tolerant Flight Control System for
Unmanned Aerial Vehicles(Mechanical Systems). Transactions of the Japan Society of Mechanical Engineers Series C 75:756,
2301-2310. [Crossref]
49. . Review 3-26. [Crossref]
50. Youmin Zhang, Jin Jiang. 2008. Bibliographical review on reconfigurable fault-tolerant control systems. Annual Reviews
in Control 32:2, 229-252. [Crossref]
51. Jovan Boskovic, Nathan Knoebel, Raman Mehra, Irene Gregory. An Integrated Approach to Damage Accommodation in
Flight Control . [Citation] [PDF] [PDF Plus]
52. Jovan Boskovic, Joseph Jackson, Nhan Nguyen, Raman Mehra. Multiple Model-Based Adaptive Fault-Tolerant Control
of Delta Clipper Experimental (DC-X) Planetary Lander . [Citation] [PDF] [PDF Plus]
53. N. Lechevin, C.A. Rabbath, M. Shanmugavel, A. Tsourdos, B.A. White. An integrated decision, control and fault detection
scheme for cooperating unmanned vehicle formations 1997-2002. [Crossref]
54. Jovan D. Boskovic, Nathan Knoebel, Raman K. Mehra. An initial study of a combined robust and adaptive control approach
to guaranteed performance flight control 5150-5155. [Crossref]
55. Nicolas Léchevin, Camille Alain Rabbath. 2007. Robust Decentralized Fault Detection in Leader-to-Follower Formations
of Uncertain, Linearly Parameterized Systems. Journal of Guidance, Control, and Dynamics 30:5, 1528-1536. [Citation]
[PDF] [PDF Plus]
Downloaded by INDIAN INSTITUTE OF TECHNOLOGY on April 3, 2019 | http://arc.aiaa.org | DOI: 10.2514/1.11272

56. Jovan Boskovic, Joshua Redding, Raman Mehra. Integrated Health Monitoring and Adaptive Reconfigurable Control .
[Citation] [PDF] [PDF Plus]
57. Jovan D. Boskovic, Joshua Redding, Raman K. Mehra. Stable Adaptive Reconfigurable Flight Control with Self-Diagnostics
5765-5770. [Crossref]
58. Jovan D. Boskovic, Joshua Redding, Raman K. Mehra. Robust Fault-Tolerant Flight Control using a New Failure
Parameterization 5753-5758. [Crossref]
59. Jovan D. Boskovic, Ravi Prasanth, Raman K. Mehra. 2007. Retrofit Fault-Tolerant Flight Control Design Under Control
Effector Damage. Journal of Guidance, Control, and Dynamics 30:3, 703-712. [Citation] [PDF] [PDF Plus]
60. J.D. Boskovic, R.K. Mehra. A multiple model-based decentralized system for accommodation of failures in second-order
flight control actuators 6 pp.. [Crossref]

You might also like