Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Accommodation of Control Actuator Failures in Morphing Aircraft∗

Jovan D. Bošković†and Joshua Redding‡


Scientific Systems Company, Inc.

In this paper we describe a new algorithm for accommodation of failures in morphing flight control actu-
ators. The algorithm is well suited for the case of continuous actuators that can assume arbitrary positions
between the saturation limits. The algorithm is based on introducing a high-order space of actual control in-
puts (ACI) and a low-order space of virtual control inputs (VCI), and singular value decomposition (SVD) of
the problem into the control-law design in the VCI space, and control allocation which depends on the mapping
between the VCI and ACI. Based on this decomposition, a new parameterization of failures is derived in such
a way that the failures in the ACI space are reflected as pure parametric uncertainty in the VCI. On the basis
of this parameterization, an adaptive observer and stable adaptive laws for estimation of uncertain parameters
are designed, and a proof of stability is given. Performance of the overall system is evaluated on a simulation of
F/A-18 aircraft where four standard flight control surfaces each are replaced by 25 morphing surfaces.

I. Introduction

Recent studies1–7 have shown that adapting wing geometry to meet flight load requirements yields optimal perfor-
mance during disparate mission roles and allows multi-role missions that are not feasible with fixed geometry aircraft.
Operational aircraft vary wing sweep and extend flaps to achieve aerodynamic performance improvements during flight,
such as lowering stall speeds, reducing drag during cruise, or increasing the maneuverability. Such shape changes are
obtained with conventional actuators and mechanisms. Though they produce large scale flow changes, they provide
little control over the flow and, hence, limit achievable agility and vehicle stability. Morphing aircraft structures that
utilize a large number of small shape change effectors provide the degrees of freedom required for large scale flow con-
trol and the ability to meet dynamic flight load requirements. Many issues related to actuation, flight control, control
allocation and failure accommodation that are well-understood in traditional flight control, are at present unresolved in
the context of morphing.4, 8–11 The purpose of this paper is to address some of the control design issues by designing
an integrated adaptive fault-tolerant control and control allocation system (AFTC-CAS) for morphing aircraft.
In the context of control of aircraft with morphing surfaces, control design challenges include addressing the fol-
lowing questions:
1. How should the large number of actuators be used in a most effective way? In other words, what are the possible
control strategies that can improve the performance of the aircraft with respect to criteria such as fuel consumption,
robustness, and fault tolerance.
2. What are the hardware issues (weight, electronics, actuator design)? Do the benefits of using morphing actuators
outweigh the increased complexity of the actuation and control systems?
3. How should such aircraft be controlled (grouping, control allocation)? What is the best strategy, and what are the
optimization criteria?
4. How can we achieve fault tolerance since an increased number of actuators also increases the probability of failures?
In this paper our focus is on the problem of fault-tolerant control design for aircraft with morphing actuators.

This research was supported by Air Force under Contract No. FA8650-05-M-3555 to Scientific Systems Company, Inc.
† Principal Research Engineer & Intelligent & Autonomous Control Systems Group Leader, 500 W. Cummings Park, Suite 3000, Woburn, MA,
01801, AIAA Senior Member, jovan@ssci.com
‡ PhD candidate, Massachussets Institute of Technology, on leave from SSCI, AIAA Member, jredding@mit.edu

1 of 15

American Institute of Aeronautics and Astronautics


To the best of our knowledge, there are only a few results on accommodation of actuator failures in morphing
actuators, for instance the results by Chen et al.12 However, their approach is limited due to the following: (i) It is well
suited only for actuators that operate in an “on-off” mode, and (ii) The effects of different actuators are assumed to be
linearly related. Both assumptions are restrictive and it is not clear if the approach could be extended to a more general
case.
In this report we describe a new algorithm for accommodation of failures in morphing flight control actuators. The
algorithm is well suited for the case of continuous actuators that can assume arbitrary positions between the saturation
limits. The algorithm is based on introducing a high-order space of actual control inputs (ACI) and a low-order space
of virtual control inputs (VCI), and singular value decomposition (SVD) of the problem into the control-law design
in the VCI space, and control allocation which depends on the mapping between the VCI and ACI. Based on this
decomposition, a new parameterization of failures is derived in such a way that the failures in the ACI space are
reflected as pure parametric uncertainty in the VCI. On the basis of this parameterization, an adaptive observer and
stable adaptive laws for estimation of uncertain parameters are designed, and a proof of stability is given. Performance
of the overall system is evaluated on a simulation of F/A-18 aircraft where four standard flight control surfaces each
are replaced by 25 morphing surfaces.

II. Actuator Failures

Flight control actuator failures can be broadly divided into two categories: (i) Failures that result in a total loss of
effectiveness of the control effector; and (ii) Failures that cause partial loss of effectiveness. The former includes Lock-
In-Place (LIP), float, and Hard-Over Failure (HOF), while the latter is referred to as the Loss-Of-Effectiveness (LOE)
type of failure. In the case of LIP failures the actuator “freezes” at a certain condition and does not respond to subse-
quent commands. HOF is characterized by the actuator moving to and remaining at the upper or lower position limit
regardless of the subsequent commands. The speed with which the effector moves to its position limit is constrained
by the actuator rate limit. Float failure occurs when the actuator contributes zero moment to the control authority. Loss
of effectiveness is characterized by lowering the actuator gain with respect to its nominal value. Different types of
actuator failures are shown in Figure 1. If the actuators are characterized by fast response, different types of actuator

new trim
u(t) u(t) point
uT
_ 0
u

tF Time t Time
F

(a) Lock−in−Place (b) Float

u(t) .
arctan(u max) uc (t) u(t)

tF Time

(c) Hard−over (d) Loss of Effectiveness

Figure 1. Types of Control Actuator Failures

2 of 15

American Institute of Aeronautics and Astronautics


failures can be parameterized as follows:




 uci (t), ki (t) = 1, for all t ≥ 0 No-Failure Case









 ki (t)uci (t), 0 < ǫi ≤ ki (t) < 1, for all t ≥ tFi Loss of Effectiveness




ui (t) = 


 0, ki (t) = 1, for all t ≥ tFi Float Type of Failure




uci (tFi ), ki (t) = 1, for all t ≥ tFi Lock-in-Place Failure










 (ui )min or (ui )max , ki (t) = 1, for all t ≥ tFi Hard-Over Failure

where tFi denotes the time instant of failure of the ith actuator, ki denotes its effectiveness coefficient such that ki ∈
[ǫki , 1], and ǫki > 0 denotes its minimum effectiveness.
Uncertainty associated with failures is due to: (i) Unknown time of failure tFi ; (ii) Unknown LOE coefficient ki ;
and (iii) Unknown value at which the control effector locks. This is illustrated in Figure 2. We will consider these types
of failures in the context of morphing wing design.

η
ηmax uc u
ηmin
~ arctan(k)

Uncertainty: Effector Angle η Uncertainty: Actuator Gain k

ηmin < η < ηmax 0< k <1

Figure 2. Uncertainty due to control effector failures

III. Problem Statement


One way of achieving a morphing design is to replace standard control surfaces with a large number of smaller
surfaces that collectively achieve similar control effect. We will focus on this type of design and develop the control
reconfiguration algorithms for the case of failures in the morphing actuators.
Let the plant dynamics be described by:
ẋ = Ax + Bv v, (1)
where an n-vector x and an m-vector v denote respectively the state and control input vectors, A ∈ IRn×n , and Bv ∈ IRn×m ,
where m >> n. It is seen that the actuator dynamics is neglected for the moment.
In this paper the focus will be on a class of failure scenarios that satisfy the following assumption:
Assumption 1:
(a) Up to m − n actuators can undergo total LOE failure.
(b) All actuators can undergo partial LOE failure.
(c) rank(Bv BTv ) = n.
Reference Model: The reference model is chosen in the form:
ẋ∗ = Am x∗ + Bmr, (2)
∗ n×n
where x is the state of the reference model, Am ∈ IR is an asymptotically stable matrix, and r is an n-vector of
bounded piece-wise continuous reference inputs (commands).
Control Objective: The objective is to design a control law uc (t) such that the error x(t) − x∗ (t) tends to zero asymp-
totically even in the presence of different flight control actuator failures.

3 of 15

American Institute of Aeronautics and Astronautics


IV. Baseline Control Design

The control law can be designed in the v space as follows:

v = BTv (Bv BTv )−1 (−Ax + Am x + Bm r), (3)

and can be readily shown to achieve the objective in the no-failure case.
However, as we will show in the following section, control design in the space of virtual control inputs is more
convenient for reconfigurable control design in the presence of failures.
Virtual Control Inputs: Such control inputs are defined in the n-space as:

Bu u = Bvv, (4)

where u ∈ IRn and Bu ∈ IRn×n .


The relationship between the virtual and actual control inputs can be elegantlly established using the SVD of the
matrix Bv, i.e.

Bv = UDM, (5)

where U is a n × n matrix and M is an n × m matrix such that UU T = MM T = In×n , and D is a diagonal n × n matrix.
We now choose:

Bu = UD, (6)

so that Bu u = Bv v = Bu Mv. Since it can be readily shown that, if Bv is full rank, so is Bu , it follows that u = Mv. One
possible solution for v from the last equation is:

v = M T (MM T )−1 u = M T u, (7)

since MM T = I. This is an important relationship since it avoids matrix inversion.


The system model in the (x, u) space is now:

ẋ = Ax + Bu u, (8)

and the corresponding control law is:

u = BTu (Bu BTu )−1 (−Ax + Am x + Bm r). (9)

It is seen that this control law needs to be passed through the control allocation algorithm (7) to affect the actual
plant.
The control law (9) will be used as a basis for reconfigurable control design discussed in the following section.

V. Adaptive Reconfigurable Control Design for Morphing Actuators

In this section we derive the adaptive reconfigurable control design for morphing actuators. The design is based on
the SVD described in the previous section, and the new failure parameterization described below.

A. Failure Parameterization

We will first derive the failure parameterization and corresponding on-line Failure Detection, Identification and Re-
configuration (FDIR) algorithms for the case of insignificant actuator dynamics, followed by the discussion of the case
when the actuator dynamics needs to be taken into account.

4 of 15

American Institute of Aeronautics and Astronautics


As shown in our previous work,13, 14 in the case of insignificant actuator dynamics, total and partial LOE failures
can be conveniently parameterized as follows:

v = KΣvc + (I − Σ)v̄, (10)

where v = [v1 v2 ... vm ]T is the vector of actuator outputs, vc = [vc1 vc2 ... vcm ]T is the vector of controller outputs,
K = diag[k1 k2 ... km ]T , ki ∈ [0, 1], Σ = diag[σ1 σ2 ... σm ]T ,

 1, t < tFi ,


σi =  (11)
 0, t ≥ tFi ,

where tFi denotes the time of failure of the ith actuator, v̄ = diag[v̄1 v̄2 ... v̄m ]T , and v̄i is the value at which the ith
actuator locks at tFi .
Hence, in the no-failure case, v = vc . In the case of lock-in-place failure of the ith actuator, vi (t) = v̄i , ∀t ≥ tFi , and
in the case of partial LOE failure, vi (t) = ki vci ∀t ≥ tFi , where ki < 1.
As shown previously,14 this parameterization can be effectively used to design reconfigurable controllers that assure
closed-loop stability in the presence of multiple simultaneous or sequential failures. This is accomplished by on-line
estimation of the elements of Σ and K, and the use of these estimates in the control law. However, since there are
3m unknown parameters (elements of Σ, K and v̄) and at least 3m parameter estimates need to be adjusted on-line, the
applicability of this approach is limited in the systems where m is large.
We note that the parameterization (10) describes the failures in the Actual Control Input (ACI) space, which is of
the order m, where m is very large. Hence the previous results14 are not directly applicable to the case of morphing
actuators.
For this reason we propose a new approach based on projecting the parameterization (10) from the ACI space to
the space of Virtual Control Inputs (VCI space), where failures can be described by a substantially smaller number of
parameters. The way this is accomplished is described below.
The New Failure Parameterization: When operating in the ACI space, the state equation is of the form:

ẋ = Ax + Bv[KΣvc + (I − Σ)v̄]. (12)

Since vc = M T uc and Bv = Bu M, we obtain the state equation in the VCI space in the form:

ẋ = Ax + Bu M[KΣM T uc + (I − Σ)v̄]. (13)

We now define the n × n matrix Θ∗ and an n-vector ξ∗ as:

Θ∗ = MKΣM T , (14)

ξ = M(I − Σ)v̄. (15)

Hence:

ẋ = Ax + Bu [Θ∗ uc + ξ∗ ], (16)

where Θ∗ and ξ∗ are uncertain.


It is important to note that all failures in the high-order ACI space can be represented as pure parametric uncertainty
in the lower-order VCI space.
The proposed failure parameterization has the following properties:
(a) In the no-failure case, Θ∗ = In×n (since, in that case, Σ = K = I, and since MM T = I), and ξ∗ = 0.
(b) Based on the properties of the SVD, 0 ≤ θii∗ ≤ 1 for i = 1, 2, .., m, and −1 ≤ θi∗j ≤ 1 for i = 1, 2, ..., m, j = 1, 2, ..., m,
and i , j.
(c) Let kvk ≤ vmax . It follows that kv̄k ≤ vmax . Hence kξ∗ k ≤ kMkvmax .

5 of 15

American Institute of Aeronautics and Astronautics


Hence the representation of failures in the 3m parameter space is reduced to the representation in the n(n + 1) space,
which substantially simlifies the FDIR design as discussed in the following sections. To get a feel for the reduction in
the number of parameters, we note that the maximum number of the virtual control inputs in the case of flight control
is n = 6, which results in n(n + 1) = 42 parameters, and is independent of the number of actual actuators. On the other
hand, for instance in the case of F/A-18 there are 10 control surfaces. If each is replaced by only 10 morphing surfaces,
we will have that m = 100, and the number of failure-related parameters is already 3m = 300.

B. Higher-order Actuator Dynamics

Our focus will be on the case when the dynamics of the ith actuator in the no-failure case can be expressed by a transfer
function Wi (s) of order nai whose relative degree (the difference between the number of poles and number of zeros) is
also nai :
λi1
Wi (s) = . (17)
snai + snai −1 λnai −1 + ... + λi2 s + λi1

This results in a state-space model of the form:

v̇i1 = vi2 , (18)


v̇i2 = vi3 , (19)
..
. (20)
nai
X
v̇inai = − λi j vi j − λi1 (vic − vi1 ), (21)
j=2

where λi j > 0 are chosen so that the resulting transfer function is stable, vi1 is the actuator output, and vic is its input
(i.e. the output of the controller).
Assertion: Let the values of λi j be such that the subsystem:

v̇i2 = vi3 , (22)


..
. (23)
nai
X
v̇inai = − λi j vi j , (24)
j=2

is asymptotically stable and has the property that the convergence of vnai (t) to zero is fast. Then the actuator dynamics
in the presence of total and partial LOE failures can be described by:

vi = Wi (s)[ki σi vic + (1 − σi )v̄i ], (25)

modulo exponentially decaying initial conditions.


Proof: The state space model that corresponds to the above transfer function representation is of the form:

v̇i1 = vi2 , (26)


v̇i2 = vi3 , (27)
..
. (28)
nai
X
v̇inai = − λi j vi j − λi1 σi (ki vic − vi1 ). (29)
j=2

Hence, when σi = 0, vnai (t) → 0 and the rate of convergence is fast, so that the model accurately represents lock-in-
place failures. When ki = k̄i < 1, the unit step response can be readily shown to converge to k̄i . It can be concluded that
the above model is a good representation of both total and partial LOE failures. 

6 of 15

American Institute of Aeronautics and Astronautics


The representation (25) can be also expressed in the form:

vi = ki σi vicF + (1 − σi )v̄i · 1F , (30)

modulo exponentially decaying initial conditions, where

vicF = Wi (s) · vic , (31)


F
1 = Wi (s) · 1. (32)

Let vcF = [v1c


F F
v2c F
... vmc ]. Then the state equation of the system is of the form:

ẋ = Ax + Bv [KΣvcF + (I − Σ)v̄ · 1F ], (33)


= Ax + Bu M[KΣM T ucF + (I − Σ)v̄ · 1F ], (34)
= Ax + Bu [Θ∗ ucF + ξ∗ · 1F ], (35)

where uF = MvcF .
We note that uF = uc and 1F = 1 in the case of zero-order actuator dynamics. Also, if λi1 >> λi j for i = 1, 2, ..., m
and j = 2, 3, ..., nai, then the actuator dynamics in many cases can be neglected. In this paper we will assume that the
latter is true.

C. Parameter Estimation

The new parameterization proposed in the previous section enables a straightforward parameter estimation scheme to
be designed. This is discussed below.
Adaptive Observer: Let ê = x̂ − x denote the state estimation error. The adaptive observer is chosen in the form:

x̂˙ = Ax + Bu [Θ̂uc + ξ̂] + Λê, (36)

where Λ is asymptotically stable.


Adaptive Laws: Let θ̄ = {Θ : 0 ≤ θii ≤ 1, i = 1, 2, ..., 6, −1 ≤ θi j ≤ 1, i = 1, 2, ..., 6, j = 1, 2, ..., 6, i , j}, and
ξ̄ = {ξ : kξk ≤ kMkvmax } denote the sets of parameter values. Adaptive laws are now chosen in the form:
˙
Θ̂ = Projθ̄ {−Γθ BTu êuTc }, (37)
ξ̂˙ = Projξ̄ {−Γξ B ê}, T
(38)

where Γθ = ΓTθ > 0 and Γξ = ΓTξ > 0. In the above expressions Proj{·} {·} denotes the Projection Operator whose role is
to project the parameter estimates to the corresponding set of parameter values. More details on the properties of the
Projection Operator can be found in the literature.13
Theorem 1: Adaptive laws (37), (38) assure that all the signals in the system (16), (36) are bounded, and, in addition,
ê ∈ L∞ ∩ L2 .
Proof: To prove the theorem, we first introduce: Φθ = Θ̂ − Θ∗ and φξ = ξ̂ − ξ∗ . We now subtract the plant equation (16)
from (36) to obtain the following error model:

ê˙ = −λI + Bu [Φθ uc + φξ ], (39)

where, to simplify the analysis, we have chosen Λ = −λI.


˙ and φ̇ = ξ̂.
We next note that, since Θ∗ and ξ∗ are constant, Φ̇θ = Θ̂ ˙
ξ

Now the following Lyapunov function candidate is chosen:


1 T
V(ê, Φθ , φξ ) = [ê ê + trace(ΦTθ Γ−1 T −1
θ Φθ ) + φξ Γξ φξ ], (40)
2

7 of 15

American Institute of Aeronautics and Astronautics


where Γθ = ΓTθ > 0 and Γξ = ΓTξ > 0 denote the adaptive gain matrices.
Its first derivative along the solutions of the system yields:
V̇ ≤ −λkêk2 ≤ 0, (41)
T T
where the following was used: (i) The fact that trace(ab ) = a b, and (ii) Properties of the Projection Operator that,
given the error equation ė = −λe + ωT φ and adaptive law φ̇ = Projθ̄ {−eω}, we have that φφ̇ ≤ −eφω for all values of
the arguments.13
From (41) it follows that ê is bounded (Θ̂ and ξ̂ are already bounded by the choice of the adaptive law). We now
integrate (41) from 0 to ∞ to obtain:
Z ∞
V(0) − V(∞) ≤ λ kê(τ)k2 dτ. (42)
0

Since all the signals are bounded, so is V(∞). Hence ê ∈ L ∩ L2 . 
2
While the above theorem guarantees that ê is bounded and belongs to L , this does not imply that either x or x̂ is
bounded. This, and the proof of the convergence of the tracking error to zero is discussed in the following section.
Significant Actuator Dynamics: In the case of significant actuator dynamics, the adaptive laws are chosen in the
form:
˙ = Proj {−Γ BT ê(uF )T },
Θ̂ (43)
θ̄ θ u c
˙ξ̂ = Proj {−Γ BT ê · 1F }. (44)
ξ̄ ξ

D. Adaptive Reconfigurable Control

The adaptive reconfigurable controller is now chosen in the form:


uc = (BuΘ̂)T [Bu Θ̂(Bu Θ̂)T ]−1 (η − Bu ξ̂), (45)
where
η = −Ax + Am x + Bm r. (46)

Theorem 2: The control law (45), where the parameters are adjusted using (37) and (38), assures that the tracking
error e = x − xm is bounded, and, limt→∞ e(t) = 0.
Proof: We first substitute the control law (45) into the observer (36), and after some straigtforward manipulations
obtain:
x̂˙ = Am x̂ + Bm r − (Am + λI)ê. (47)
Let em = x̂ − xm . From (2) and the above equation we now have:
ėm = Am em − (Am + λI)ê. (48)
Since the boundedness of ê can be established independently of em , using BIBO stability argument it follows that em
is bounded. Since xm is bounded, this implies that x̂ is bounded as well. This, based on the boundedness of ê, now
implies that x is bounded as well.
Hence all the signals in the system are bounded. We now use Barbalat’s lemma36 to show that limt→∞ ê(t) = 0.
This implies, based on BIBO stability arguments for (48), that limt→∞ em (t) = 0. Since e = em − ê, it now follows that
limt→∞ e(t) = 0. 
From close inspection of the control law it is seen that there is a possibility for the matrix Bu Θ̂(t) to lose rank at
some time instant, invalidating the control law. One possibility for alleviating this problem is to use the One-Step-
Ahead-Controller (OSAC),38 which is of the form:
uc = [(Bu Θ̂)T QBu Θ̂ + R]−1 (Bu Θ̂)T Q(η − Bu ξ̂), (49)
T T
where Q = Q ≥ 0, R = R > 0, and η is given by (46). It is seen that the control law is well defined even when the
matrix Bu Θ̂(t) = 0 for some t. This control law can be shown to result in bounded tracking error.

8 of 15

American Institute of Aeronautics and Astronautics


E. Extensions to Nonlinear Dynamics Case

In this section our focus is on the following nonlinear model of aircraft dynamics:

ẋ = f (x) + gv (x)v, (50)

where x : IR+ → IRn denotes the state vector, f : IRn → IRn , gv : IRn → IRn×m , and v : IR+ → IRm denotes the plant
input vector. It is assumed that the above model describes the dominant dynamics of the aircraft. In such a case, the
state variables include forward velocity, V, angle-of-attack, α, side-slip angle, β, angular velocities, p, q and r, and
attitude angles, φ, θ and ψ. In addition, it is assumed that the nonlinearities, f (x) and gv (x) are associated with nominal
aircraft dynamics (i.e. the dynamics in flight regimes in which there are no failures or damage).
The above model is subject to the following assumption:
Assumption 1:
(a) State of the system is measurable;
(b) For a closed bounded set of states S x , gv (x)gTv (x) is invertible for all x ∈ S x ;
(c) f (x) and gv (x) are sufficiently smooth functions (functionals) of their argument;
(d) m >> n; and
(e) Up to m − n actuators can undergo LIP failures, while all actuators can undergo LOE failures.
The desired dynamics of the aircraft is chosen in the form:

ẋm = Am xm + Bm r, (51)

where xm : IR+ → IRn denotes the state of the reference model, matrix Am is asymptotically stable, and r : IR+ → IRn
denotes a vector of bounded piece-wise continuous reference inputs.
VCI Definition in the Nonlinear Case: Let us assume that the control allocation algorithms is of the form:

M(x)v = u, (52)

where M(x) is full rank for all x in the set S x , and u is the vector of virtual control inputs. Hence on S x we have that:

v = M(x)T (M(x)M(x)T )−1 u. (53)

This also implies that:

vc = M(x)T (M(x)M(x)T )−1 uc . (54)

Failure Model in the Nonlinear Case: We use the same actuator failure parameterization as in the linear case:

v = KΣvc + (I − Σ)v̄. (55)

After substituting it into the plant equation one obtains:

ẋ = f (x) + gv (x)[KΣvc + (I − Σ)v̄]. (56)

From (54) it now follows:

ẋ = f (x) + gv (x)[KΣM(x)(M(x)M(x)T )−1 uc + (I − Σ)v̄] (57)


= f (x) + Θ(x)uc + ξ(x), (58)

where Θ(x) = gv (x)KΣM(x)(M(x)M(x)T )−1 and ξ(x) = gv (x)(I − Σ)v̄.


It is seen that in this case Θ and ξ are functions of x and linear adaptive control theory can be no longer applied. For
this reason we use neural networks to identify these nonlinearities on-line, and use these estimates in the control law
to achieve effective failure accommodation. The approach that we use is similar others,15–17 and is briefly described
below.

9 of 15

American Institute of Aeronautics and Astronautics


Estimation of the Effect of Failures using Neural Networks: We first express the term Θ(x)uc in the form:
n
X
Θ(x)uc = θi (x)uci , (59)
i=1

where θi (x) is the ith column of Θ(x). Hence:


n
X
ẋ = f (x) + θi (x)uci + xi(x). (60)
i=1

To design an observer for the unknown quantities, θi (x) and ξ(x) are first approximated by neural networks over a
compact set of states as:

θi (x)  Fi (x)wi + ǫi (x), i = 1, 2, ..., n, (61)


ξ(x)  Fn+1 (x)wn+1 + ǫn+1 (x), (62)

where wi are vectors of the neural network weights, and ǫi (x) is the approximation error. The functionals Fi (x) are
obtained by using either Radial Basis Function Neural Networks (RBF-NN) of Multi-layer NN (MNN) with adjustable
weights in the output layer. The parameterization of MNN that is linear in the network weights is obtained by fixing
the weights in the hidden layer(s) and adjusting only the weights in the output layer. Similarly, linear parameterization
in the case of RBF-NN is obtained by fixing the centers and widths of the radial basis functions.
The resulting plant equation is:

ẋ = f (x) + F(x)w + ǫ(x), (63)


Pn+1
T
where F(x) = [F1T (x) F2T (x) ... Fn+1 (x)]T , w = [wT1 wT2 ... wTn+1 ]T , and ǫ = i=1 ǫi (x).
The observer is now introduced in the form:

x̂˙ = f (x) + F(x)ŵ + Λ( x̂ − x), (64)

where Λ is asymptotically stable.


Let φ = ŵ − w, and ê = x̂ − x. Then the error model is of the form:

ê˙ = Λê + Ω(x, θ̂)φ − ǫ(x). (65)

Let P = PT > 0 be a solution of the Lyapunov matrix equation ΛT P + PΛ = −Q, where Q = QT > 0. Now the
adjustment laws for φ̂ is of the form:

φ̂˙ = ŵ˙ = −ΓF T (x)Pê − γ0 kêkθ̂, (66)

whereΓ = ΓT > 0 and γ0 > 0 denote adaptive gains. It can be readily shown that the above adjustment laws assure the
boundedness of the tracking error.
The Adaptive Fault-Tolerant Controller is of the form of OSAC:

uc = [Θ̂T QΘ̂ + R]−1 Θ̂T Q(η − ξ̂), (67)

where Q = QT ≥ 0, R = RT > 0, and η is given by:

η = − f (x) + Am x + Bm r.

It is seen that the control law is well defined even when the matrix Θ̂(t) = 0 for some t.
The main challenges related to the neural network design include: (i) Implementing the multiple neural networks
resulting from the above parameterization in an effective way; (ii) Finding the most suitable neural network structure
(RBF-NN or MNN, number of layers and nodes, adjustment gains); and (iii) Assuring acceptable response.

10 of 15

American Institute of Aeronautics and Astronautics


VI. Simulations
The simulations were carried out on a semi-nonlinear simulation of F/A-18 dynamics. The F/A-18 model was
modified so that the left and right TEFs and AILs were replaced by 25 small surfaces each. Toghether with the
standard control inputs, this results in a total of 100 control effectors. Four test cases are included. A brief description
of the simulations and failure cases is noted below.
Test Vehicle & Flight Condition: As a test vehicle we chose F/A-18C/D at the flight condition 0.7M and altitude
of 20,000 feet.
F/A-18 Simulation: For this flight condition we obtained a simulation of F/A-18 dynamics from Boeing with linear A
and B matrices, nonlinear kinematics and gravity effects, actuator dynamics, and position and rate limits ont he control
effectors.
Control Effectors: The F/A-18 has the following left and right control effectors: Leading-Edge Flaps (LEF), Trailing-
Edge Flaps (TEF), Ailerons (AIL), Stabilators (STAB), Rudders (RUD), and engines (ENG). Left and right TEFs and
AILs are replaced by 25 small surfaces each. Toghether with the standard control inputs, this results in a total of 100
control effectors.
Test Maneuver: Test maneuver is a 30◦ roll doublet while maintaining course.
Failure Cases: Failure cases include:
1. Failure Case 1: at t = 2 sec, failure of TEFs 1-13 and 26-37, and Ailerons: 1-13 and 26-37 (symmetric failure), and
2. Failure Case 2: at t = 2 sec, failure of TEFs 1-22 and Ailerons: 1-23 (asymmetric failure).
Simulation Cases: Simulation cases are listed in Table 1.
Case Controller Failure
1 Baseline VCI No
2 Reconfigurable VCI No
3 Baseline VCI Yes (Failure Case 1)
4 Reconfigurable VCI Yes (Failure Case 1)
5 Reconfigurable VCI Yes (Failure Case 2)

Table 1. List of Simulation Cases

Simulation Case 1: The Baseline VCI is designed along the lines presented in the previous sections. The matrix
Bv from the F/A-18 simulation was decomposed using SVD and matrices Bu and M were found. The controller was
designed using the matrix Bu , and the control inputs were allocated in the ACI space using matrix M. The resulting
response for the 30o roll doublet are given in Figure 3.
Simulation Case 2: The adaptive fault-tolerant reconfigurable (AFTR) controller space was designed in the VCI
space. The objective of this simulation is to demonstrate that the AFTR controller achieves performance similar to that
obtained using the Baseline VCI controller, i.e. that it does not destabilize the closed loop system. The response with
the AFTR controller is given in Figure 4, and is seen to be very close to that obtained in Figure 3.
Simulation Case 3: In this case the aircraft is controlled by the Baseline VCI controller, and the failure scenario 1 was
applied at t = 2, and the resulting response is shown if Figure 5. It is seen that, without failure accommodation, the
response is unacceptable and the system becomes unstable.
Simulation Case 4: The response of the system under the same failure scenario but with the AFTR controller is given
in Fig. 6. The AFTR controller is seen to achieve effective failure compensation and acceptable performance of the
closed-loop system.
Simulation Case 5: It can be readily shown that, in the case of failure scenario 2, the closed-loop system with the
Baseline VCI becomes unstable. The response with the AFTR controller is shown in Fig. 7. The AFTR controller is
again seen to achieve effective failure compensation and acceptable performance of the closed-loop system. We note
that, since the failure in this case is more severe than in the Failure Case 2, the control effort needed to compensate for
the failure is larger than in the latter case.

11 of 15

American Institute of Aeronautics and Astronautics


VII. Conclusions and Future Work

In this paper we describe a new algorithm for accommodation of failures in morphing flight control actuators. The
algorithm is well suited for the case of actuators that can assume arbitrary positions between the saturation limits. The
algorithm is based on introducing a high-order space of actual control inputs (ACI) and a low-order space of virtual
control inputs (VCI), and SVD of the problem into the control-law design in the VCI space, and control allocation which
depends on the mapping between the VCI and ACI. Based on this decomposition, a new parameterization of failures is
derived in such a way that the failures in the ACI space are reflected as pure parametric uncertainty in the VCI. On the
basis of this parameterization, an adaptive observer and stable adaptive laws for estimation of uncertain parameters are
designed, and a proof of stability os given. Performance of the overall system is evaluated on a simulation of F/A-18
aircraft where four standard flight control surfaces each are replaced by 25 morphing surfaces.
Future work will include the extensions of the algorithm to the case of nonlinear aircraft dynamics, and performance
evaluation using high-fidelity simulations that include both rigid-body dynamics and aeroservoelastic effects.

References
1 J. Bowman, B. Sanders and T. Weisshaar, “Evaluating the Impact of Morphing Technologies on Aircraft Performance”, AIAA-2002-1631,
2002.
2 C. Cesnik, H. Last and C. Martin, “A Framework for Morphing Capability Assessment”, AIAA-2004-1654, Proc. 45th

AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics and Materials Conf., April, 2004.


3 S. Joshi, Z. Tidwell, W. Crossley and S. Ramakrishnan, “Comparison of Morphing Wing Strategies Based Upon Aircraft Performance

Impacts”, AIAA-2004-1722, Proc. 45th AIAA/ASME/AHS/ASC Structures,Structural Dynamics and Materials Conf., April 2004.
4 J. Kudva, C. Martin, L. Scherer, A. Jardine, A. McGowan, R. Lake, G. Sendeckyj and B. Sanders, “Overview of DARPA/USAF/NASA

Smart Wing Program”, Proc. SPIE Smart struc. and Materials Conf., 1999.
5
B. Prock, T. Weisshaar and W. Crossley, “Morphing Airfoil Shape Change Optimization With Minimum Actuator Energy As An Objective”,
AIAA-2002-5401, Proc. 9th AIAA/ISSMO Symposium on Multidisciplinary Analysis and Optimization, Atlanta, GA, September 2002.
6
B. Sanders, F. Eastep and E. Forster, “Aerodynamic and Aeroelastic Characteristics of Wings with Conformal Control Surfaces for Morphing
Aircraft”, Journal of Aircraft, Vol. 40, No. 1, January-February, 2003.
7
T. Weisshaar, “Aeroelastic tailoring for energy efficient morphing aircraft - finding the right stuff”, ICASE/NASA LaRC, October, 2001. See
also Professor Weisshaar’s keynote addresses at SPIE (March 2003) and AIAA SDM Conference (April, 2003).
8 DARPA Morphing Aircraft Structures (MAS) program, http://www.darpa.mil/dso/thrust/matdev/mas.html
9 R. Wlezien, G. Horner, A. McGowan, S. Padula, M. Scott, R. Silcox, J. Simpson, “The aircraft morphing program”, AIAA 1998-1927, Proc.

AIAA/ASME/ASCE/AHS/ASC Structures, Struc. Dyn. and Materials Conf., 1998.


10 G. Tao, S. Chen, J. Fei, S. M. Joshi, “An adaptive actuator failure compensation scheme for controlling a morphing aircraft model”, Proc.

Conf. Decision and Control, 2003.


11 D. Raney, R. Montgomery, L. Green and M. Park, “Flight control using distributed shape change effector arrays”, Proc. 41st AIAA/ASME/AHS

Structures, Structural Dynamics and Materials Conf., AIAA Paper 2000-1560, Atlanta, GA, April 2000.
12 S. H. Chen, G. Tao, J. T. Fei and S. M. Joshi, “Adaptive compensation of morphing actuator failures”, in Proceedings of the 2004 American

Control Conference, pp. 1805–1810, Boston, MA, 2004.


13 J. D. Bošković and R. K. Mehra, “A Multiple Model Adaptive Flight Control Scheme for Accommodation of Actuator Failures”, AIAA

Journal of Guidance, Control & Dynamics, Vol 25, No. 4, pp. 712-724, July-August, 2002.
14 J. D. Bošković, S.-H. Yu, and R. K. Mehra, “A Stable Scheme for Automatic Control Reconfiguration in the Presence of Actuator Failures”,

in Proceedings of the 1998 American Control Conference, Vol. 4, pp. 2455-2459, Philadelphia, PA, June 24-26, 1998.
15 K. S. Narendra and K. Parthasarathy, “Identification and Control of Dynamical Systems Using Neural Networks”, IEEE Transactions on

Neural Networks, Vol. 1, pp. 4-27, 1990.


16 K. S. Narendra and S. Mukhopadhyay, “Adaptive Control of Nonlinear Multivariable Systems Using Neural Networks”, Neural Networks,

Vol. 7, No. 5, pp. 737-752, 1994.


17 K. S. Narendra and S. Mukhopadhyay, “Intelligent Control Using Neural Networks”, in M. M. Gupta and K. Sinha (Eds.), Intelligent Control

Systems–Theory and Applications, IEEE Press, NY, 1996.


18 A. Barron, “Universal Approximation Bounds for Superpositions of a Sigmoidal Function”, IEEE Transactions on Information Theory, Vol.

39, No. 3, pp. 930-945, May 1993.


19 M. Bolender and D. Doman, “Non-linear Control Allocation Using Piecewise Linear Functions”, Journal of Guidance, Control and Dynam-

ics, Vol 27, No. 6, Nov-Dec, 2004.


20 J. D. Bošković, S. E. Bergstrom, R. K. Mehra, J. Urnes, Sr., M. Hood, and Y. Lin, “Fast on-Line Actuator Reconfiguration Enabling (FLARE)

System”, in Proceedings of the 2005 AIAA Guidance, Navigation and Control Conference, San Francisco, CA, August 15-18, 2005.
21 J. D. Bošković, R. Prasanth and R. K. Mehra, “Retrofit Reconfigurable Flight Control under Control Effector Damage”, AIAA Journal of

Guidance, Control & Dynamics, Vol. 30, No. 3, pp. 703-712, May-June 2007.
22 J. D. Bošković and R. K. Mehra, “Robust Integrated Flight Control Design Under Failures, Damage and State-Dependent Disturbances”,

AIAA Journal of Guidance, Control & Dynamics, Vol. 28, No. 5, pp. 902-917, September-October 2005.

12 of 15

American Institute of Aeronautics and Astronautics


23 J. D. Bošković and R. K. Mehra, “Control Allocation in Overactuated Aircraft under Position and Rate Limiting”, in Proceedings of the

2002 American Control Conference, Anchorage, Alaska, May 8-10, 2002.


24 J. D. Bošković and R. K. Mehra, “An Adaptive Scheme for Compensation of Loss of Effectiveness of Flight Control Effectors”, in Proceed-

ings of the 40th IEEE Conference on Decision and Control, Orlando, FL, December 2001.
25 J. D. Bošković, S.-M. Li and R. K. Mehra, “On-Line Failure Detection and Identification (FDI) and Adaptive Reconfigurable Control (ARC)

in Aerospace Applications”, in Proceedings of the 2001 American Control Conference, pp. 2625-2626, Arlington, VA, June 25-27, 2001.
26 J. D. Bošković and R. K. Mehra, “Intelligent Adaptive Control of a Tailless Advanced Fighter Aircraft under Wing Damage”, AIAA Journal

of Guidance, Control & Dynamics, Vol. 23, No. 5, pp. 876-884, September-October 2000.
27 J. D. Bošković, S.-M. Li and R. K. Mehra, “Reconfigurable Flight Control Design Using Multiple Switching Controllers and On-line

Estimation of Damage-Related Parameters”, in Proceedings of the 2000 Conference on Control Applications, Anchorage, Alaska, September 2000.
28 J. D. Bošković, “A Stable Neural Network-Based Adaptive Control Scheme for a Class of Nonlinear Plants”, in Proceeding of the 1997

Conference on Decision and Control, Vol. 1, pp. 472-476, San Diego, CA, December 1997.
29 J. D. Bošković and K. S. Narendra, “Comparison of Linear, Nonlinear and Neural Network-Based Adaptive Controllers for a Class of

Fed-Batch Processes”, Automatica, Vol. 31, No. 6, pp. 817-840, 1995.


30 J. Brinker and K. Wise, “Flight Testing of a Reconfigurable Flight Control Law on the X-36 Tailless Fighter Aircraft”, Proc.1998 AIAA GNC

Conference, Paper No. AIAA-2000-3941, Denver, CO, August 2000.


31 J. Brinker and K. Wise, “Reconfigurable Flight Control of a Tailless Advanced Fighter Aircraft”, Proc.1998 AIAA GNC Conference, Vol. 1,

pp. 75-87, Boston, MA, Aug. 1998.


32 A. Calise, S. Lee and M. Sharma, “Direct Adaptive Reconfigurable Control of a Tailless Fighter Aircraft”, Proc. 1998 AIAA GNC Conference,

Vol. 1, pp. 88-97, Boston, MA, Aug. 1998.


33 D. Enns, “Control Allocation Approaches”, in Proceedings of the 1998 AIAA Guidance, Navigation and Control (GNC) Conference, Boston,

MA, August 2000.


34
S. Haykin, Neural Networks – A Comprehensive Foundation, Macmillan College Publishing Company, 1994.
35 K. Hornik, M. Stinchcombe and H. White, “Multilayer Feedforward Networks are Universal Approximators”, Neural Networks, Vol. 2, pp.

359-366, 1989.
36 K. S. Narendra and A. M. Annaswamy, Stable Adaptive Systems, Prentice Hall Inc., Englewood Cliffs, New Jersey, 1988.
37 E. W. Pendleton, D. Bessette, P. B. Field, G. D. Miller and K. E. Griffin, “Active Aeroelastic Wing Flight Research Program: Technical

Program and Model Analytical Development”, Journal of Aircraft, Vol. 37, No. 4, 2000, pp. 554-561.
38 R. K. Prasanth and J. D. Bošković, “Stability Analysis and Design of Optimal One Step Ahead Controllers”, in Proceedings of the 2005

AIAA Guidance, Navigation and Control Conference, San Francisco, CA, August 15-18, 2005.
39 A. Vemuri and M. Polycarpou, “Robust nonlinear fault diagnosis in input-output systems”, Int. J. Control, Vol.68, No.2, 1997.

5
*
β,β*[deg]

β(t), β (t)
V(t), V*(t)
V,V*[ft/s]

0
725.8

725.7 −5
0 5 10 15 0 5 10 15
0.2 50
p,p*[deg/s]

p(t), p*(t)
q,q [deg/s]

q(t), q*(t)
0 0
*

−0.2 −50
0 5 10 15 0 5 10 15
1
2.8
r,r*[deg/s]

θ(t), θ*(t) r(t), r*(t)


θ,θ*[deg]

0
2.6
−1
0 5 10 15 0 5 10 15
50
2.8 *
φ,φ [deg]

α(t), α*(t)
α,α*[deg]

φ(t), φ (t)
0
*

2.6
−50
0 5 10 15 0 5 10 15
2001 1
ψ(t), ψ*(t)
ψ,ψ*[deg]

h(t), h*(t)
h,h [ft]

2000 0
*

1999 −1
0 5 10 15 0 5 10 15

3.6
LL,LR [deg]

u1(t), u2(t)
3.4

3.2
0 5 10 15
TL,TR [deg]

4 u3(t)−u52(t)

3
0 5 10 15
EL,ER [deg] RL,RR [deg] SL,SR [deg] AL,AR [deg]

2
u53(t)−u77(t)
0

−2
0 5 10 15
5
u103(t), u104(t)
0

−5
0 5 10 15
1
u (t), u (t)
105 106
0

−1
0 5 10 15
70
u107(t), u108(t)
65
60
0 5 10 15

Figure 3. Case 1. Response with the Baseline VCI controller in the no-failure case.
13 of 15

American Institute of Aeronautics and Astronautics


β(t), β*(t)
2

β,β [deg]
V(t), V*(t)

V,V [ft/s]
0
725.8

*
*
725.7 −2
0 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 14 16 18 20
0.2 50

p,p*[deg/s]
p(t), p*(t)

q,q [deg/s]
*
q(t), q (t)
0 0

* −0.2 −50
0 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 14 16 18 20
1
2.8

r,r*[deg/s]
* r(t), r*(t)
θ,θ [deg]

θ(t), θ (t)
0
*

2.6
−1
0 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 14 16 18 20
50
2.8 *

φ,φ*[deg]
*
α,α [deg]

α(t), α (t) φ(t), φ (t)


0
*

2.6
−50
0 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 14 16 18 20
2001 1
*

ψ,ψ [deg]
* ψ(t), ψ (t)
h(t), h (t)
h,h [ft]

2000 0
*

*
1999 −1
0 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 14 16 18 20

LL,LR [deg] 3.6


u1(t), u2(t)
3.4

3.2
0 2 4 6 8 10 12 14 16 18 20
TL,TR [deg]

4 u3(t)−u52(t)

3
0 2 4 6 8 10 12 14 16 18 20
EL,ER [deg] RL,RR [deg] SL,SR [deg] AL,AR [deg]

2
u53(t)−u77(t)
0

−2
0 2 4 6 8 10 12 14 16 18 20
2
u103(t), u104(t)
0

−2
0 2 4 6 8 10 12 14 16 18 20
1
u105(t), u106(t)
0

−1
0 2 4 6 8 10 12 14 16 18 20
70
u107(t), u108(t)
65
60
0 2 4 6 8 10 12 14 16 18 20

Figure 4. Case 2. Response with the Adaptive Fault-Tolerant Reconfigurable (AFTR) VCI controller in the no-failure case.
10
725.8
β,β*[deg]
V,V [ft/s]

*
5 β(t), β*(t)
725.78 V(t), V (t)
*

0
725.76 0 2 4 6 8 10 12 14 16 18 20
0 2 4 6 8 10 12 14 16 18 20 50
0.1
p,p*[deg/s]

p(t), p*(t)
q,q [deg/s]

0
0
*

q(t), q*(t)
−50
−0.1 0 2 4 6 8 10 12 14 16 18 20
0 2 4 6 8 10 12 14 16 18 20 2
3.5
r,r*[deg/s]

r(t), r*(t)
θ,θ [deg]

*
θ(t), θ (t) 0
3
*

2.5 −2
0 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 14 16 18 20
2.8 50 *
φ(t), φ (t)
φ,φ*[deg]
α,α [deg]

* 0
α(t), α (t)
*

2.6 −50
0 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 14 16 18 20
2050 10
*
ψ,ψ*[deg]

* ψ(t), ψ (t)
h(t), h (t)
h,h [ft]

2000 0
*

1950 −10
0 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 14 16 18 20

3.6
LL,LR [deg]

3.4
u1(t), u2(t)
3.2
0 2 4 6 8 10 12 14 16 18 20
TL,TR [deg]

u3(t)−u17(t), u28(t)−u42(t)
4

2 u18(t)−u27(t), u43(t)−u52(t)
0 2 4 6 8 10 12 14 16 18 20
AL,AR [deg]

2 u68(t)−u77(t), u93−u102
0
u53(t)−u67(t), u78−u92
−2
0 2 4 6 8 10 12 14 16 18 20
SL,SR [deg]

10
u (t)
103
0
u104(t)
−10
0 2 4 6 8 10 12 14 16 18 20
EL,ER [deg] RL,RR [deg]

0.5 u105(t), u106(t)

0
0 2 4 6 8 10 12 14 16 18 20
70
u107(t)
65
60 u (t)
108
0 2 4 6 8 10 12 14 16 18 20

Figure 5. Case 3. Response with the Baseline VCI controller in the failure case 1.

14 of 15

American Institute of Aeronautics and Astronautics


5
*

β,β [deg]
β(t), β (t)
V(t), V*(t)

V,V [ft/s]
0
725.8

*
*
725.7 −5
0 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 14 16 18 20
0.2 50

p,p*[deg/s]
p(t), p*(t)

q,q [deg/s]
*
q(t), q (t)
0 0

*
−0.2 −50
0 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 14 16 18 20
1
2.8

r,r*[deg/s]
* r(t), r*(t)
θ,θ [deg]
θ(t), θ (t)
0
*

2.6
−1
0 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 14 16 18 20
50
2.8 *

φ,φ*[deg]
*
α,α [deg]

α(t), α (t) φ(t), φ (t)


0
*

2.6
−50
0 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 14 16 18 20
2002 1
*

ψ,ψ [deg]
* ψ(t), ψ (t)
h(t), h (t)
h,h [ft]

2000 0
*

*
1998 −1
0 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 14 16 18 20

3.6
LL,LR [deg] u1(t), u2(t)
3.4

3.2
0 2 4 6 8 10 12 14 16 18 20
TL,TR [deg]

u3(t)−u17(t), u28(t)−u42(t)
4

2 u18(t)−u27(t), u43(t)−u52(t)
0 2 4 6 8 10 12 14 16 18 20
EL,ER [deg] RL,RR [deg] SL,SR [deg] AL,AR [deg]

2 u53(t)−u67(t), u78−u92
0
u68(t)−u77(t), u93−u102
−2
0 2 4 6 8 10 12 14 16 18 20
2
u103(t), u104(t)
0

−2
0 2 4 6 8 10 12 14 16 18 20
1
u105(t), u106(t)
0

−1
0 2 4 6 8 10 12 14 16 18 20

64 u107(t), u108(t)

62
0 2 4 6 8 10 12 14 16 18 20

Figure 6. Case 4. Response with the Adaptive Fault-Tolerant Reconfigurable (AFTR) VCI controller in the failure case 1.

2
726 *
β,β [deg]

* β(t), β (t)
V,V [ft/s]

V(t), V (t)
0
725.8
*
*

725.6 −2
0 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 14 16 18 20
2 50
p,p*[deg/s]
q,q*[deg/s]

*
q(t), q (t) p(t), p*(t)
0 0

−2 −50
0 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 14 16 18 20
4 1
r,r*[deg/s]

r(t), r*(t)
θ,θ [deg]

*
3 θ(t), θ (t) 0
*

2 −1
0 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 14 16 18 20
4 50
*
φ,φ*[deg]

α(t), α (t) *
α,α [deg]

3 φ(t), φ (t)
0
*

2
1 −50
0 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 14 16 18 20
1
2050 * *
ψ,ψ [deg]

h(t), h (t) ψ(t), ψ (t)


h,h [ft]

0
*

2000

1950 −1
0 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 14 16 18 20
LL,LR [deg]

3.6
u1(t), u2(t)
3.4
3.2
0 2 4 6 8 10 12 14 16 18 20
6
TL,TR [deg]

u26−u27(t), u28(t)−u52(t)
4
u3(t)−u25(t)
2
0 2 4 6 8 10 12 14 16 18 20
EL,ER [deg] RL,RR [deg] SL,SR [deg] AL,AR [deg]

1
u53(t)−u75(t)
0
u (t)−u (t), u −u
76 77 78 102
−1
0 2 4 6 8 10 12 14 16 18 20
2
u103(t), u104(t)
0

−2
0 2 4 6 8 10 12 14 16 18 20
1
u105(t), u106(t)
0

−1
0 2 4 6 8 10 12 14 16 18 20

64 u107(t), u108(t)

62
0 2 4 6 8 10 12 14 16 18 20

Figure 7. Case 5. Response with the Adaptive Fault-Tolerant Reconfigurable (AFTR) VCI controller in the failure case 2.

15 of 15

American Institute of Aeronautics and Astronautics

You might also like