You are on page 1of 15

Automatica 43 (2007) 1869 – 1883

www.elsevier.com/locate/automatica

Adaptive actuator failure compensation for nonlinear MIMO systems with an


aircraft control application夡
Xidong Tang a , Gang Tao b,∗ , Suresh M. Joshi c
a GM R&D and Planning, Mail Code 480-106-390, Warren, MI 48090, USA
b Department of ECE, University of Virginia, Charlottesville, VA 22903, USA
c Mail Stop 308, NASA Langley Research Center, Hampton, VA 23681, USA

Received 18 March 2005; received in revised form 29 November 2006; accepted 14 March 2007
Available online 15 August 2007

Abstract
A direct adaptive approach is developed for control of a class of multi-input multi-output (MIMO) nonlinear systems in the presence
of uncertain failures of redundant actuators. An adaptive failure compensation controller is designed which is capable of accommodating
uncertainties in actuator failure time instants, values and patterns. A realistic situation is studied with fixed grouping of actuators and proportional
actuation within actuator groups. The adaptive control system is analyzed, to show its desired stability and asymptotic tracking properties in
the presence of actuator failure uncertainties. As an application, such an adaptive controller is used for actuator failure compensation of a twin
otter aircraft longitudinal model, with design conditions verified and control structure and adaptive laws developed for a nonlinear aircraft
dynamic model. The effectiveness of adaptive failure compensation is demonstrated by simulation results.
䉷 2007 Elsevier Ltd. All rights reserved.

Keywords: Actuator failure; Adaptive control; Backstepping; Multivariable systems; Nonlinear control

1. Introduction Maybeck, 1999), fault detection and diagnosis-based designs


(Demetriou & Polycarpou, 1998; Vemuri & Polycarpou, 1997;
Actuator failures can cause severe performance deteriora- Wang & Daley, 1996; Wu, Zhang, & Zhou, 2000; Zhang,
tion of control systems, or even system instability leading to Parisini, & Polycarpou, 2004), neural network-based designs
catastrophic accidents. Compensation of uncertain actuator fail- (Calise, Lee, & Sharma, 2001; Diao & Passino, 2001; Wise,
ures has been an important and challenging research problem. Brinker, Calise, Enns, & Elgersma, 1999), and sliding mode
Remarkable progresses have been made in the area of actua- control-based designs (Corradini & Orlando, 2003). Recently
tor failure accommodation control with various effective de- adaptive actuator failure compensation control schemes have
sign methods developed such as adaptive designs (Ahmed-Zaid, been developed based on a direct adaptive control approach
Ioannou, Gousman, & Rooney, 1991; Bodson & Groszkiewicz, for compensation of unknown actuator failures (Chen, Tao,
1997; Boskovic, Yu, & Mehra, 1998b), multiple-model designs & Joshi, 2004; Mirkin & Gutman, 2005; Tang, Tao, & Joshi,
(Boskovic & Mehra, 1999; Boskovic, Yu, & Mehra, 1998a; 2002; Tao, Chen, Tang, & Joshi, 2004). Unlike designs based
on some other approaches, direct adaptive failure compensa-
tion control designs are model-based, that is, the nominal sys-
夡 This paper was not presented at any IFAC meeting. This paper was tem structural information is incorporated into adaptive failure
recommended for publication in revised form by Associate Editor Sam Ge compensation designs, and more importantly, they are able to
under the direction of Editor Miroslav Krstic. This research was supported ensure desired system stability and asymptotic tracking proper-
by the NASA Langley Research Center Research Grant NCC-1342 and the ties, despite the presence of uncertain failures. For many appli-
NSF Research Grant ECS0601475.
∗ Corresponding author. cations, asymptotic tracking is a desirable property of control
E-mail addresses: xidong.tang@gm.com (X. Tang), gt9s@virginia.edu systems in which certain system structural information is often
(G. Tao), s.m.joshi@larc.nasa.gov (S.M. Joshi). available.
0005-1098/$ - see front matter 䉷 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.automatica.2007.03.019
1870 X. Tang et al. / Automatica 43 (2007) 1869 – 1883

A direct adaptive control scheme is specifically designed to resulted without enough elevator power, which could be made
handle uncertainties in both system dynamics and actuator fail- uncatastrophic if the landing is controlled well). The work of
ures that can occur during system operation. Such failures are this paper is aimed at the design and analysis of an adaptive
often uncertain in time, value, and pattern, that is, it is un- control scheme for the considered system, which is capable of
known when, how much, and which actuators fail. Compared utilizing the remaining actuation power to make a desired sys-
with other designs, direct adaptive failure compensation con- tem motion, despite the unknown system failures.
trol designs have simpler controller structures: only one con- Systems failures introduce additional and large systems un-
troller is used to accommodate the system dynamics change certainties. For example, when some actuators fail, the system
caused by actuator failures. They adaptively adjust controller structure from the active inputs (unfailed actuators) to the out-
parameters, using adaptive laws driven by system response er- put experiences significant changes, and so do the system pa-
rors, to achieve desired system performance. Incorporated with rameters. Furthermore, failures cause additional disturbances
system structural information, they can be designed to handle to systems, which influence a system’s behavior at different
large uncertainties of parameters of the actuator failures and actuation locations. Compensation of such disturbances alone
the controlled system as well, to ensure, in addition to stabil- can be a nontrivial problem (i.e., certain matching conditions
ity, the asymptotic tracking of a reference signal by the system have to be established by a proper controller design). In this
signal. sense, system failures do not just cause system gain changes,
In our previous work on direct adaptive control for actu- and they lead to system uncertainties which need to be treated
ator failure compensation, an adaptive scheme for compen- by new adaptive control designs and analyzed by new adaptive
sation of actuator failures was developed for linear MIMO control theory, for which there are open issues. Furthermore,
systems (Chen et al., 2004), and the actuator failure compensa- the robustness of adaptive failure compensation designs with
tion problem was studied for single-output nonlinear systems in respect to system modeling errors such as unmodeled dynam-
a parametric-strict-feedback canonical form (Tang et al., 2002). ics and disturbances is also an important issue. This issue can
In this paper, we will address the problem for nonlinear MIMO be addressed but it will not be done in this paper. It has been
systems with application to aircraft flight control. We will con- well-understood that adaptive controllers have certain inher-
sider the control of a class of nonlinear MIMO systems with ent robustness properties (Tao, 2003), and with robust adaptive
uncertain parameters, in the presence of uncertain failures of laws, they are robust with respect to bounded disturbances and
redundant actuators in each group of related physical character- unmodeled dynamics (Ioannou & Sun, 1996).
istics. We will develop a direct adaptive failure compensation The paper is organized as follows. The control problem for-
control scheme to meet desired system performance specifica- mulation is given in Section 2. The design and analysis of a
tions as discussed in the above paragraph. multivariable adaptive actuator failure compensation scheme
System redundancy is widely employed in the design of mod- are given in Section 3. A detailed study (analysis, design and
ern control systems such as aircraft flight control system, space simulation) of application to aircraft flight control in the pres-
structure systems and other performance-critical systems. The ence of actuator failures is given in Section 4 for a nonlinear
existence of redundancy brings new challenges for feedback twin otter aircraft longitudinal model.
control designs, especially, when the system failure pattern is
unknown. A desirable controller should be able to handle all 2. Problem statement
possible failure patterns and values, with all possible degrees of
redundancy. In this sense, redundancy is a part of the problem, Consider the following nonlinear system with m inputs and
not a part of the proposed approach for solving the problem. q outputs (m > q):
Effective solutions to this problem are crucial for many ap-
plications including systems with MEM devices, aircraft flight ẋ = f0 (x) + f (x) + g(x)u,
control, intelligent robots, space structures, especially, for their y = h(x), (1)
autonomous operations.
We will consider the actuator failure compensation problem where x ∈ R n , y = [y1 , y2 , . . . , yq ]T ∈ R q , u = [u1 , u2 , . . . ,
for up to m − qa (q qa m) actuator failures, where m (the um ]T ∈ R m whose components (actuators) may fail during
number of actuators) and q (the number of system outputs) are operation, f0 (x) ∈ R n , f (x) = [f1 (x), f2 (x), . . . , fl (x)]T ∈
determined by a specific application; in particular, m is not nec- R n×l , g(x) = [g1 (x), g2 (x), . . . , gm (x)] ∈ R n×m , h(x) ∈ R q ,
essarily large, while qa is not necessarily small. For example, an and  ∈ R l is an unknown constant parameter vector.
aircraft may only have 2–4 ailerons of which only one of them Actuator failure model: The actuator failures considered in
may fail during a flight. How to effectively (adaptively) utilize this paper are modeled as
the remaining (unfailed) actuators for a given control task (e.g.,
aircraft safe landing) is the key open issue to be solved, even if uj (t) = ūj (t), t tj , j ∈ {1, 2, . . . , m}, (2)
full actuator authority for a given task can be recovered using
where the failure time instant tj is unknown, and
fewer actuators. As for the case when full actuator authority
(actuation power) cannot be recovered using fewer actuators, 
s
desirable system performance cannot be expected due to lack ūj (t) = ūj k j k (t), t tj , j ∈ {1, 2, . . . , m} (3)
of actuator authority (e.g., a likely unsafe landing would be k=1
X. Tang et al. / Automatica 43 (2007) 1869 – 1883 1871

for some unknown constants ūj k and unknown pattern (index) vector u = [u1 , u2 , . . . , um ]T ∈ R m can be expressed as
j, and known signals j k (t), k = 1, 2, . . . , s.
u = v(t) + (ū(t) − v(t)), (4)
Remark 1. The failure model (3) is in a completely
where v(t) = [v1 (t), v2 (t), . . . , vm (t)]T is the applied control
parametrized form. This is only an ideal model for the study in
input vector to be generated from a feedback law, ū(t) =
this paper, as the knowledge of the functions j k (t) may not be
[ū1 (t), ū2 (t), . . . , ūm (t)]T is the failure vector, and
available for some applications. In such cases, approximations
of j k (t) can be employed to achieve approximate compensa-  = diag{1 , 2 , . . . , m },
tion of actuator failures. Some commonly used approximation 
1 if the ith actuator fails at t: ui = ūi (t),
methods in the literature employ a set of basis functions such i = i (t) = (5)
as Taylor series and neural networks to approximate the con- 0 otherwise.
sidered unknown nonlinear functions. With approximation for
This (t) is the actuator failure pattern indicator (its specific
failure functions, there will be an approximation error and
value represents a specific failure pattern).
the magnitude of the error is mainly determined by the num-
Suppose that at time t there are p actuator failures in the
ber of basis functions aj k (t) used in approximation. A such
a system, that is, uj = ūj (t), j = j1 , j2 , . . . , jp , for some
approximation expression is ūj (t) = sk=1 ūaj k aj k (t) + j (t), {j1 , j2 , . . . , jp } ⊂ {1, 2, . . . , m}, leading to a certain fail-
for some unknown constants ūaj k , known basis signals aj k (t), ure pattern  with jk = 1, k = 1, 2, . . . , p, and jk = 0,
k = 1, 2, . . . , s a , and some unknown and unparametrizable k  = 1, 2, . . . , p. With (4), system (1) can be described as
error signal j (t). Such an approximation error j (t) is
bounded for a bounded set of j k (t) and can be made small ẋ = f0 (x) + f (x) + g(x)ū(t) + g(x)(I − )v,
by increasing the number s a of basis functions aj k (t). This y = h(x). (6)
approximation error will appear in the closed-loop control
system and can be handled using robust adaptive control the- It can be seen that when actuator failures take place, not only
ory, by modifying the adaptive parameter update laws with the corresponding applied control inputs will be stopped, but
an additional signal which guarantees the control system’s also uncertainties will be brought into the system, especially,
robustness with respect to bounded error signals. As a result, the controlled system changes its structure when the failure
the closed-loop stability can still be ensured and the tracking pattern changes. A desirable control design should be able to
error can be made small but asymptotic tracking cannot be handle all possible failure patterns of interest. Next, we develop
ensured. a direct adaptive control scheme which can accommodate the
unknown actuator failures, without using explicit failure detec-
Control objective: The control objective is to design an adap- tion. The essence of such a direct adaptive failure compensa-
tive controller for system (1) with actuator failures (3) with tion controller is that it adaptively searches some new direc-
unknown failure time instants tj , failure pattern j and fail- tion for the vector of parameter estimates whenever there is a
ure parameters ūkj , to guarantee that the remaining actuators new failure pattern, such that after a transient response the re-
can still ensure boundedness of the closed-loop signals and tuned controller ensures desired system performance, despite
asymptotic tracking of a given reference output signal yr (t) = the presence of failure uncertainties.
[yr1 (t), yr2 (t), . . . , yrq (t)]T by the plant output y(t), in the
presence of up to m − qa (q qa m) actuator failures. 3. Adaptive failure compensation design
The value of qa depends on the designer’s choice for a spe-
cific application; in particular, qa is not necessarily small (for A controller which can achieve stability and asymptotic
example, qa = m − 1 could be a choice, so that the problem tracking for one failure pattern may not be suitable for another
is for up to m − qa = 1 actuator failure, that is, either there is failure pattern. For each chosen control design, there is a set of
no failure or there is one failure of the total m actuators, but failure patterns which can be compensated under certain sys-
it is not known which one fails or maybe no one fails). In this tem conditions. Our goal is to enlarge the set of compensable
paper, we derive the solution for the case of qa = q (in fact, actuator failures while relaxing control system design condi-
the case with up to m − q failures includes the case with up to tions. A failure pattern is called a compensable failure pattern
m − qa failures for q < qa m). of a certain control design if the associated actuator failures
On the other hand, for output tracking, it is necessary to can be compensated (that is, closed-loop stability and asymp-
have at least q inputs to be functional (unfailed) at any time, totic tracking are ensured despite actuator failure uncertainties)
in order to have enough independent control inputs to drive by the control design. The compensable failure pattern set of
the q independent outputs to track some arbitrary reference a control design for a system is the set of all failure patterns
output signals. Hence, to meet the control objective, there compensable by the control design.
can only be at most m − q actuator failures allowed in the We note that the compensable failure pattern set which is
system. the largest set of failures that can be compensated by a control
As derived in Boskovic et al. (1998b) and Tao et al. (2004), design is determined by both the controlled system and the
in the presence of possible actuator failures, the system input control design as indicated by the results of Tang et al. (2002).
1872 X. Tang et al. / Automatica 43 (2007) 1869 – 1883

If we study systems in canonical forms such as the parametric- grouping scheme denoted as
strict-feedback form with the consideration of some specific
G : {{j1,1 , j1,2 , . . . , j1,d1 }, . . . , {jk,1 , jk,2 , . . . , jk,dk }
actuation schemes, sufficient conditions can be derived for the
existence of an adaptive compensation scheme for system (1) , . . . , {jq,1 , jq,2 , . . . , jq,dq }} (8)
with actuator failures on a compensable failure pattern set. We
such that there are dk inputs uj in the same group k,
start with the following definition.
j = jk,1 , jk,2 , . . . , jk,dk , {jk,1 , jk,2 , . . . , jk,dk } ⊂ {1, 2, . . . , m}.
This actuator grouping scheme has its meaningful practical
Definition 1 (Marino and Tomei, 1995). The control char-
feature: actuators of the same or similar physical character-
acteristic indices of system (1) are the set of integers
istics are naturally in one group, while actuators of different
{1 , 2 , . . . , q }, 1i n, i = 1, 2, . . . , q, such that:
physical characteristics usually should not be in one group.
Then, a proportional actuation scheme is employed to ma-
Lgj Lkf0 hi (x) = 0, j = 1, 2, . . . , m, k = 0, 1, . . . , i − 2, nipulate the design redundancy, that is,
 −1
Lgj Lf0i hi (x)  = 0 for some j = ji ∈ {1, 2, . . . , m}, (7) vj =bj (x)wk for j =jk,1 , jk,2 , . . . , jk,dk , k=1, 2, . . . , q, (9)
where wk is the nominal control for group k to be de-
where Lf p is the Lie derivative of a scalar function p(x) along signed, k = 1, 2, . . . , q, and bj (x) is some design function
the vector field f (x) = [f1 , f2 , . . . , fn ]T , defined as Lf p = of certain physical meaning for the actuators in each group,
(jp/jx)f = (jp/jx1 )f1 + · · · + (jp/jxn )fn . j =jk,1 , jk,2 , . . . , jk,dk for each actuator group (j =1, 2, . . . , m,
together for all actuator groups).
Remark 2. The existence of well-defined control characteristic With this actuator grouping and actuation scheme, we group
indices for a system guarantees that each output is connected to the actuators in advance of control design so that the group-
some inputs, which is a basic requirement for output tracking ing scheme G is fixed, for which some design conditions on
of a reference so that the actuator failure compensation may be the system (1), the grouping scheme (8), and the design func-
achieved by a control design implemented with known param- tions bj (x) in (9) are given next. Such conditions are needed
eters and failures. However, since the inputs may be dependent for a nominal (nonadaptive) design with known actuator fail-
on each other, a condition on the independence of the control ures, and, as we will show, they are also sufficient for a stable
inputs is needed for achieving output tracking, as specified in adaptive design when the knowledge of actuator failures is un-
Theorem 1 (as the condition (i)). known.
With the well-defined control characteristic indices {1 , 2 ,
The following assumption imposes a triangularity condi- . . . , q } (see Definition 1 and Assumption (A1)), the fixed
tion on fk (x), k = 1, 2, . . . , l, for characterizing the system grouping scheme G and the actuation scheme (9), system (1)
in the parametric-strict-feedback form which can be controlled with p actuator failures in a failure pattern : {j1 , j2 , . . . , jp }
by the backstepping technique (Krstić, Kanellakopoulos, & can be transformed into a canonical form which consists of
Kokotović, 1995). a tracking dynamic subsystem and a zero dynamic subsys-
tem via a diffeomorphism [T , T ]T = [TcT (x), TzT (x)]T ∈ R n ,
Assumption (A1). The system (1) has well-defined control where
characteristic indices {1 , 2 , . . . , q } in a domain U defined
 −1
by h: U ⊂ R n → V ⊂ R q such that yr (t) ∈ V , for ∀t > 0. The Tc (x) = [h1 (x), Lf0 (x) h1 (x), . . . , Lf01(x) h1 (x), h2 (x)
functions fk (x) satisfy the strict triangularity condition  −1  −1
, . . . , Lf02(x) h2 (x), . . . , Lf0q(x) hq (x)]T , (10)
 − +r −1
d(Lfk Lfri0−1 hi ) ∈ span{d(h1 ), . . . , d(Lf01 i i h1 ),  = [1,1 , 1,2 , . . . , 1,1 , 2,1 , . . . , q,q ]T ∈ R 1 +2 +···+q ,
 −i +ri −1
d(h2 ), . . . , d(Lf02 h2 ) and  ∈ R n−(1 +2 +···+q ) .
The tracking dynamics are given by
 −i +ri −1
, . . . , d(Lf0j hj ), . . . , d(hq )
˙ 1,1 = 1,2 + 1,1 (1,1 , 2,1 , . . . , 2,2 −1 +1 , . . . , q,1
T
 − +ri −1
, . . . , d(Lf0q i hq )} , . . . , q,q −1 +1 ),

for ri = 1, . . . , i − 1, k = 1, 2, . . . , l, i = 1, 2, . . . , q, where ˙ 1,2 = 1,3 + 1,2 (1,1 , 1,2 , 2,1 , . . . , 2,2 −1 +2
T

d(f ) is the differential of f defined on a vector field X, that , . . . , q,1 , . . . , q,q −1 +2 )
is, d(f ) = (jf/jx) dx, where dx is the basis vector for the
cotangent bundle of X (Isidori, 1995). ..
.
Actuator grouping and actuation scheme: Based on the con- ˙ 1,1 −1 = 1,1 + 1,1 −1 (1,1 , . . . , 1,1 −1 , 2,1
T
trol characteristic indices {1 , . . . , q }, the m actuators are
grouped into q groups where q is the number of outputs, by a , . . . , 2,2 −1 , . . . , q,1 , . . . , q,q −1 ),
X. Tang et al. / Automatica 43 (2007) 1869 – 1883 1873

˙ 1,1 = 1,1 (, ) +
T
1,j (, )ūj (t) 1,1 , 1,2 , . . . , 1,1 −i +ri , 2,1 , . . . , q,q −i +ri . In the SISO
j =j1 ,...,jp case, it defines a lower triangular form.
Notice that Tc (x) is independent of parameters, actuator fail-

q
+ ¯ ures and a certain control strategy, while Tz (x) is changed
1,k (, )wk
k=1
with actuator failures and determined by the chosen actua-
tion scheme and grouping scheme. Therefore, the zero dynam-
.. ics in general are dependent on the actuator failure pattern 
. and determined by the chosen control strategy and grouping
˙ i,ri = i,ri +1 + i,ri (1,1 , 1,1 −i +ri , . . . , i,1
T
scheme.
Now we present the main result of this paper.
, . . . , i,ri , . . . , q,1 , . . . , q,q −i +ri )

.. Theorem 1. Under Assumption (A1), the grouping scheme G in


. (8) and the actuation scheme (9), system (1) is compensable for
 an actuator failure pattern set  by an adaptive control scheme,
˙ i,i = i,i (, ) +
T
i,j (, )ūj (t) if for any failure pattern  ∈ , the following conditions
j =j1 ,...,jp
hold:

q
+ ¯
i,k (, )wk (i) the following equivalent actuation matrix:
k=1 ⎡¯ ¯ 1,q ⎤
1,1 · · ·
..
. M (x) = ⎣ · · · · · · · · · ⎦ (14)
˙ q,1 = q,2 + ¯ ¯ q,q 
q,1 (1,1 , . . . , 1,1 −q +1 , 2,1 q,1 · · ·
T

, . . . , 2,2 −q +1 , . . . , q,1 ), is nonsingular in U; and


(ii) the zero dynamics (12) are input-to-state stable (ISS) in U
˙ q,2 = q,3 + q,2 (1,1 , . . . , 1,1 −q +2 , 2,1
T
for any failure pattern  ∈  with respect to the failure
signal ū(t) and  as the inputs.
, . . . , 2,2 −q +2 , . . . , q,2 )
Remark 3. Conditions (i) and (ii) above define the compens-
..
. able set  under the adaptive control scheme to be constructed
next (in the Proof of Theorem 1). Condition (i) implies
˙ q, q −1
= q,q + q,q −1 (1,1 , . . . , 1,1 −1 , 2,1
T
that for any failure pattern  ∈ , the control characteris-
, . . . , 2,2 −1 , . . . , q,1 , . . . , q,q −1 ), tic indices {1 , 2 , . . . , q } are preserved by the remaining
 working (unfailed) actuators uj , j  = j1 , j2 , . . . , jp , that is,
 −1
˙ q,q = Tq,q (, ) + q,j (, )ūj (t) Lgj Lf0i hi (x)  = 0, for some j ∈ / {j1 , j2 , . . . , jp }; otherwise
j =j1 ,...,jp there will be some rows of the matrix M (x) such as the
singularity of M (x) is resulted. Therefore, this condition is

q
+ ¯ to ensure that the control characteristic indices are invari-
q,k (, )wk (11)
k=1 ant and well-defined regardless of the actuators failing, for a
compensable failure pattern . With well-defined and invari-
and the zero dynamics are given by
ant control characteristic indices, condition (i) in fact ensures
˙ =  (, ) +  (, ) +  (, )ū(t), (12) the system controllability in the presence of actuator failures,
which is needed for achieving stability and asymptotic output
where i,k = 0 for k 0, and tracking.
i,ri = [Lrfi1 (x) hi (x), . . . , Lrfil (x) hi (x)]T , Condition (ii) means that when the grouping scheme and the
actuation scheme are chosen, the system needs to be minimum
i = 1, 2, . . . , q, ri = 1, . . . , i − 1, phase in the presence of actuator failures, for any compensable
 −1 failure pattern.
i,j = Lgj Lf i hi , i = 1, 2, . . . , q, j = j1 , j2 , . . . , jp ,
 Clearly, both conditions are needed for a desired (nominal)
¯  −1
i,k = bj Lgj Lf i hi , nonadaptive control design even if the characteristics of actu-
j ∈{j1 ,...,jp }∩{jk,1 ,...,jk,dk } ator failures are known. As we will show next, these condi-
tions are also sufficient for construction of a desired adaptive
i = 1, 2, . . . , q, k = 1, 2, . . . , q. (13) control design which is capable of handling uncertain actuator
The above tracking dynamics are in the parametric-strict- failures. 
feedback form (Krstić et al., 1995), that is, the non-
linearities i,ri and uncertain parameter vector  only Proof of Theorem 1. The proof is formulated by construc-
appear on the feedback path and depend on the states tion. For each failure pattern  ∈ , according to condition (i),
1874 X. Tang et al. / Automatica 43 (2007) 1869 – 1883

the nominal control inputs wk , k = 1, 2, . . . , q, can be de- T ˆ


i,ri = − ci,ri zi,ri − zi,ri −1 − i,ri 
coupled into q independent groups. Hence, a nonadaptive
j −i +ri −1
compensation control scheme can be developed to com- 
q  ji,ri −1 T ˆ
pensate the actuator failures. Considering the tracking dy- + (j,k+1 + j,k )
jj,k
namics in the form of (11) with p actuator failures in a j =1 k=1
compensable failure pattern , we develop a nonadaptive ji,ri −1
+ i,ri −1
compensation controller which compensates the actuator fail- jˆ
ures for output tracking by applying the backstepping method in ⎛ ⎞
Krstić et al. (1995).  q j − i +ri −1
jj,k
The adaptive controller is chosen as +⎝ zj,k+1 ⎠ 
j ˆ

j =1 k=1
⎛ ⎞
vj = bj (x)wk , j = jk,1 , jk,2 , . . . , jk,dk , k = 1, 2, . . . , q,  q j − i +ri −1
j i,ri −1
{jk,1 , jk,2 , . . . , jk,dk } ⊂ {1, 2, . . . , m}, × ⎝ i,ri − ⎠
jj,k j,k
⎡ ⎤ ⎡  (x, , ˆ ȳr1 ) − T
(t)ūI ⎤
j =1 k=1
w 1 1, 1 1
⎢w ⎥ ⎢  (x, , 
ri
⎢ 2⎥ ⎢ 2,2 ˆ ȳr2 ) − T
2 (t)ūI ⎥
⎥ +
ji,r −1 i (k)
y ,
w=⎢ ⎥
⎢ .. ⎥ = M −1
 (x) ⎢
⎢ ..
⎥,
⎥ jyri
(k−1) ri
⎣ . ⎦ ⎣ . ⎦ k=1

wq ˆ ȳrq ) −
q,q (x, , T
(t)ūI 
q
q i,ri = j,j −i +ri −1 + zi,ri 
˙ 
q
j =1
ˆ = i,i , (15) ⎛ ⎞
j −i +ri −1
i=1 
q  ji,ri −1
×⎝ i,ri − j,k
⎠ (16)
jj,k
 j =1
where ˆ is the estimate of , ȳri =
k=1
[yri , ẏri , . . . , yrii ]T ,
i = [ i,1 , i,2 , . . . , i,m ] , i = 1, 2, . . . , q,
T (t) = diag for ri = 1, 2, . . . , i and i = 1, 2, . . . , q, where i,0 = yri .
{1 (t), 2 (t), . . . , m (t)} with j (t) = [j 1 (t), j 2 (t), . . . , Suppose that there are N compensable failure patterns (n) ,
j s (t)]and ū = diag{ū1 , ū2 , . . . , ūm }, with ūj = [ūj 1 , ūj 2 , . . . , n = 1, 2, . . . , N, in the compensable failure pattern set  (as
ūj s ]T , I is a vector of dimension m whose components are a notation, we use (n) to indicate the indexed failure patterns
ones, i.e., I = [1, 1, . . . , 1]T ∈ R m .i,i and i,i are derived  ∈ ). For each (n) , n ∈ {1, 2, . . . , N}, there exists a non-
via a backstepping procedure. adaptive compensation scheme in the form of (15). We choose
For a simple presentation, assume that 1 2  · · · q . a controller structure under the grouping scheme G as
The backstepping procedure starts from 1,1 . From (11) we vj = bj (x)wk , j = jk,1 , jk,2 , . . . , jk,dk ,
see that 1,1 = h1 (x) = y1 . Define z1,1 = 1,1 − yr1 , and by
differentiating z1,1 we get ż1,1 = 1,2 + T1,1  − ẏr1 . De- {jk,1 , jk,2 , . . . , jk,dk } ∈ G, k = 1, 2, . . . , q,
⎡  (x, , ˆ ȳr1 ) − T (t)2 ⎤
fine a stabilizing function 1,1 = −c1,1 z1,1 − T1,1 ˆ + ẏr1 1,1 1
and z1,2 = 1,2 − 1,1 such that ż1,1 = −c1,1 z1,1 + z1,2 − ⎢ ˆ ⎥
⎢ 2,2 (x, , ȳr2 ) − 2 (t)2 ⎥
T
 N
T ( ˆ − ). w= −1
1,n Mn (x) ⎢ ⎢ ⎥, (17)
1,1 .. ⎥
Now we take a partial Lyapunov function V1,1 = 21 z1,1 2 n=1 ⎣ . ⎦
to evaluate the first step. Differentiate V1,1 and obtain ˆ T
q,q (x, , ȳrq ) − q (t)2
V̇1,1 = −c1,1 z1,12 +z z T ˆ
1,1 1,2 − z1,1 1,1 ( − ), where z1,1 z1,2 where Mn (x) with the subindex n indicates its dependence on
will be canceled by the next stabilizing function 1,2 for failure patterns (n) , n = 1, 2, . . . , N, 1,n and 2 satisfy the
z1,2 , and z1,1 T1,1 (ˆ − ) will be handled at the last step following matching conditions:
with the adaptive law for ˆ in (15) via a complete Lya- 1,n = 1 if actuators fails in the failure pattern (n) ,
punov function given later on. At this step, a tuning function
1,1 = z1,1  1,1 with the adaptation gain matrix  = T > 0 1,n = 0 otherwise,
is introduced rather than a adaptive law for ˆ to avoid over- 2 = ū(n) I (18)
parameterization. Next if 2 =1 , we follow a similar procedure
if actuators fail in the failure pattern (n) , n ∈ {1, 2, . . . , N}.
for z2,1 =2,1 −yr2 . Otherwise continue the above procedure for
The controller in the structure (17) is a combination of all
z1,2 = 1,2 − 1,1 .
controllers designed for a certain compensable failure pattern
The entire backstepping procedure is summarized as
(15) by introducing two constant vectors of parameters 1 =
[1,1 , 1,2 , . . . , 1,N ]T and 2 which in fact represent the fail-
zi,ri = i,ri − i,ri −1 (1,1 , . . . , 1,1 −i +ri −1 , . . . , i,1 ure pattern and failure parameters in actuator failures. Since
ˆ yri , ẏri , . . . , y (ri −1) ), both the failure pattern and failure parameters are unknown, an
, . . . , i,ri −1 , . . . , q,q −i +ri −1 , , ri adaptive compensation scheme will be developed by using ˆ 1
X. Tang et al. / Automatica 43 (2007) 1869 – 1883 1875

and ˆ 2 as the estimates of 1 and 2 to handle the uncertainties With the adaptive compensation design (19)–(20), the time-
based on the nonadaptive compensation design given in (15). ˆ ˆ 1 , ˆ 2 ) is derived as
derivative of V (x, ,
It is recalled that Tc (x) is independent of parameters, actuator
˙ˆ ˙ˆ 1
failures, the grouping scheme and the actuation scheme, which V̇ = zT ż + (ˆ − )T −1 ˆ 1 − 1 )T −1
  + ( 1 
means the backstepping design is irrelevant to actuator failures
+ (ˆ 2 − 2 )T −1 ˙ˆ 2
so that the functions i,i , i,i , i = 1, 2, . . . , q, and the up- 2 
date law for ˆ are unchanged by actuator failures. Therefore  i
q 
compared with the nonadaptive compensation design, the un- = − 2
ci,ri zi,ri
certainty caused by actuator failures in one fixed controller for i=1 ri =1
a certain failure pattern is 2 , while 1 changes value to switch ⎛ ⎞
i
the control for another failure pattern when new actuator fail- 
q 
ji,ri −1 ⎝ ˙ˆ 
q
− zi,ri − j,j ⎠
ures appear.
i=1 ri =2 jˆ j =1
The adaptive compensation scheme is given by ⎛ ⎞
˙ 
q
vj = bj (x)wk , j = jk,1 , jk,2 , . . . , jk,dk , + (ˆ − )T −1

⎝ˆ − j,j ⎠
{jk,1 , jk,2 , . . . , jk,dk } ⊂ {1, 2, . . . , m}, k = 1, 2, . . . , q, j =1

⎡ 
N
ˆ
1,1 (x, , ȳr1 ) −
T
1 (t)ˆ 2 ⎤ + (ˆ 1,n − 1,n )−1
n
 ⎢  (x, ,
ˆ ȳr2 ) − T
(t)ˆ 2 ⎥
N
⎢ 2,2 2 ⎥ n=1
 
ŵ = −1 ⎢
ˆ 1,n Mn (x) ⎢ ⎥,
.. ⎥ 
q
n=1 ⎣ . ⎦ × ˙ˆ 1,n + n ˆ , ˆ 2 )
zi,i ¯ Tn,i Mn−1 ¯ (x, ,
ˆ ȳrq ) −
q,q (x, , T
(t)ˆ 2 i=1
q  

N 
q

= ˆ , ˆ 2 )
ˆ 1,n Mn−1 (x)¯(x, , (19) + (ˆ 2 − 2 ) 2T −1
˙ˆ 2 − 2 zi,i
n=1 i=1
i

q 
with the adaptive laws = − 2
ci,ri zi,r 0 (23)
i
i=1 ri =1
˙ 
q
ˆ = i,i ,
for t in each time interval associated with each failure pattern
i=1
 ∈ . Note that the form of V (x, , ˆ ˆ 1 , ˆ 2 ) does not change

q
with actuator failures. However, when there appears one or
˙ˆ 1,n = −n ˆ , ˆ 2 ),
zi,i ¯ n,i Mn−1 ¯ (x, ,
T
more than one actuator failures, V (·) as a function of time t
i=1
has a jump with a limit value at that time instant because of

q
the change of 1 and 2 , which causes the piecewise continuity
˙ˆ 2 = 2 zi,i T
(t) i , (20) of V (·). Except for a finite number of jumping points, V (·)
i=1 is differentiable with a negative time derivative, which means
V (·) decreases with time from a jumping point to the next for
where ¯ n,i = [ ¯ n,i,1 , ¯ n,i,2 , . . . , ¯ n,i,q ]T for a compensable fail-
an unchanged failure pattern with no actuator failure during
ˆ , ˆ 2 ) is in (19), n=1, 2, . . . , N,  > 0
ure pattern (n) , ¯ (x, , n this time span. Therefore, we conclude that V (·) is piecewise
is the adaptation gain for ˆ 1,n and 2 = T2 > 0 is the adap- continuous and bounded. It follows that z ∈ L∞ , ˆ ∈ L∞ , ˆ 1 ∈
tation gain matrix for ˆ 2 . L∞ and ˆ 2 ∈ L∞ , and in turn we have  ∈ L∞ . From condition
Consider the Lyapunov function candidate (ii), it follows that  ∈ L∞ because the zero dynamics are ISS
with bounded ū(t) and  as the inputs. Hence, all closed-loop
ˆ ˆ 1 , ˆ 2 ) = 1 zT z + 1 (ˆ − )T −1 (ˆ − )
V (x, , 2 2  signals are bounded. According to (23), it can be seen that z ∈
+ 21 (ˆ 1 − 1 )T −1 L2 because there are limit actuator failures and (23) is satisfied
1 (
ˆ 1 − 1 )
for the final failure pattern. It also can be seen that ż ∈ L∞
+ 21 (ˆ 2 − 2 )T −1
2 (
ˆ 2 − 2 ), (21) because the closed-loop signals  and  and the derivatives of
the reference yr are bounded. Thus, limt→∞ z(t) = 0, which
where implies that limt→∞ (y − yr ) = 0. 
ˆ ˆ 1 , ˆ 2 ) = [z1,1 , . . . , z1, , . . . , zi,r , . . . , zq, ]T ,
z(x, , 1 i q
If zero dynamics are independent of actuator failures, the
1 = [1,1 , 1,2 , . . . , 1,N ] , T control system design and analysis can be simplified. Such a
condition on the structure of the actuation functions gj (x),
ˆ 1 = [ˆ 1,1 , ˆ 1,2 , . . . , ˆ 1,N ]T ,
j = jk,1 , jk,2 , . . . , jk,dk , for the actuators in the same group k
1 = diag{1 , 2 , . . . , N }. (22) is given next.
1876 X. Tang et al. / Automatica 43 (2007) 1869 – 1883

Proposition 1. For all dk actuators uj in the group k, j = uncertain due to system parameter and actuator failure uncer-
jk,1 , jk,2 , . . . , jk,dk , if their corresponding gj (x) satisfy the tainties and change as the actuator failure pattern changes. As
structure condition that a comparison, for the case when there is no actuator redun-
dancy (that is, m=q) and there is no actuator failure, a standard
gj (x) ∈ span{g0k (x)}, j = jk,1 , jk,2 , . . . , jk,dk , (24)
backstepping-based asymptotic output tracking control scheme
where g0k ∈ R n is a vector field, the inputs uj , j = jk,1 , for system (1) also requires the nonsingularity of an actuation
jk,2 , . . . , jk,dk , do not appear in zero dynamics. matrix and the input-to-state stability of a resulted zero dy-
namics system. For this case, there is only a unique grouping
Proof. With the geometric condition (24), the output zero- with each group containing only one actuator and the propor-
ing submanifold (Isidori, 1995) is same for all gj (x) in the tional actuation functions bj (x) are simple and can be chosen
group k, j = jk,1 , jk,2 , . . . , jk,dk , which means that there ex- as 1. Our adaptive failure compensation control scheme has
ists a diffeomorphism to eliminate all the corresponding uj in expanded the capacity of an adaptive backstepping controller
the state equations except some for the i th derivative of yi , to handle the uncertain actuator failures and associated system
i ∈ {1, 2, . . . , q}, no matter what actuation scheme is chosen uncertainties.
for those uj . It in turn implies that uj will not enter the zero
dynamics either. 
4. Application to aircraft flight control
The above condition in fact implies that all the actuators
are parallel such that the zero dynamics that do not contain In this section, the nonlinear longitudinal dynamics of a twin
any system inputs are the same for any grouping scheme. As a otter aircraft are used for the actuator failure compensation con-
result, the failures among those actuators are irrelevant to the trol study. We first determine a compensable failure pattern set
zero dynamics in the sense that neither their failure patterns to verify the design conditions, and then give a detailed analysis
will change the structure of the zero dynamics nor their failure of the aircraft model and develop an adaptive controller for the
signals will enter the zero dynamics. system which can compensate the unknown actuator failures
effectively and automatically. The main idea of the compen-
Remark 4. In our study, we have assumed that if an actuator sation control design for actuator failures is that we examine
fails at t =ti (that is, ui (t)= ūi (t) in the form (3)), it will stay in the physical property of actuators to choose an appropriate ac-
that failure status for all t ti . The same result holds for the case tuation scheme for corresponding actuators and investigate the
when a failed actuator can resume its normal operation, that stability of zero dynamics for the compensable failure pattern
is, ui (t) = vi (t), t tir , for some tir > ti , and for the case when set.
an actuator fails and recovers alternately for a finite number of Aircraft longitudinal model: The longitudinal motion dynam-
times. ics of the twin otter aircraft (Miller & William, 1999) can be
The developed adaptive compensation design is also appli- described as
cable to the case when an actuator fails and recovers alter-
Fx cos() + Fz sin()
nately infinitely many times, if the lengths of the time inter- V̇ = ,
vals on which the actuators do not change their failure status m
are not too small. Under this condition, all closed-loop sig- −Fx sin() + Fz cos()
˙ = q + ,
nals can also be ensured to be bounded, and over each interval mV
the tracking errors converge to residual sets whose sizes are ˙ = q,
small enough if the lengths of those time intervals are large M
enough. q̇ = , (25)
Iy
In summary, the adaptive actuator failure compensation con-
trol scheme is based on an actuator grouping and actuation where V is the velocity,  is the attack angle,  is the pitch
scheme to utilize actuator redundancy and to handle actuator angle and q is the pitch rate, m is the mass, Iy is the moment
failure uncertainties. Sufficient design conditions are given in of inertia, and
Theorem 1 as the nonsingularity of an equivalent actuation ma-
trix and the input-to-state stability of a resulted zero dynamics Fx = q̄SC x (, q, e1 , e2 ) + T1 cos 1 + T2 cos 2 − mg sin(),
system, which ensure the existence of a stable adaptive control
scheme for handling uncertain actuator failures in a compens- Fz = q̄SC z (, q, e1 , e2 ) + T1 sin 1 + T2 sin 2 + mg cos(),
able failure pattern set, in addition to system parameter uncer-
tainties. Such conditions are also needed even for the nominal M = q̄cSC m (, q, e1 , e2 ), (26)
(nonadaptive) control design for the case when failure charac-
teristics and system parameters are known, using a standard for which q̄ = 21 V 2 is the dynamic pressure,  is the air density,
backstepping technique. Under the same conditions, an adap- S is the wing area, c is the mean chord, and T1 and T2 are
tive scheme has been derived to be able to desirably update independent thrusts with corresponding thrust misalignments
the estimates of the nominal controller parameters which are 1 and 2 .
X. Tang et al. / Automatica 43 (2007) 1869 – 1883 1877

The functions Cx , Cz and Cm are of the polynomial form and , b1 , b2 , c1 , c2 , p1 , a1 , a2 , 1 , 2 , d1 , d2 are unknown
constant parameters while p0 in fact is the gravity constant
Cx = Cx1  + Cx2 2 + Cx3 + Cx4 (d1 e1 + d2 e2 ), which is known.
Cz = Cz1  + Cz2 2 + Cz3 + Cz4 (d1 e1 + d2 e2 ) + Cz5 q, Control objective: The control objective is to design an adap-
tive scheme for the nonlinear aircraft plant (28) to use the el-
Cm = Cm1  + Cm2 2 + Cm3 + Cm4 (d1 e1 + d2 e2 ) evator angles and thrusts as controls such that the pitch angle
x3 and the velocity x1 track corresponding reference signals
+ Cm5 q, (27) generated from two reference systems, while up to two control
actuators may fail during operation.
where e1 and e2 are the elevator angles of an augmented two-
Compensable failure pattern set: According to the control
piece elevator used as two actuators for our failure compensa-
objective, x3 and x1 are chosen as the output y = [x3 , x1 ]T .
tion study.
Based on Definition 1, we have 1 = 2 and 2 = 1. Examine
Remark 5. We note that the elevator force in this twin otter condition (i) in Theorem 1 and find that the following failure
aircraft model is linear in the elevator angle e although non- patterns:
linear in V, the velocity. Since V is almost same for all elevator
segments, the elevator segments act on the system through the diag{1, 0, 0, 0}, diag{0, 1, 0, 0}, diag{0, 0, 1, 0},
same nonlinear function with different parameters, i.e., propor- diag{0, 0, 0, 1}, diag{0, 0, 0, 0}, (30)
tional to each other. In other words, all elevator segments are
diag{1, 0, 1, 0}, diag{0, 1, 1, 0}, diag{1, 0, 0, 1},
same in the physical meaning, and the difference is the coeffi-
cient for each segment. diag{0, 1, 0, 1} (31)

State space description: Choosing V, ,  and q as the states satisfy the conditions.
x1 , x2 , x3 and x4 , and e1 , e2 , T1 , T2 as the inputs u1 , u2 , u3
and u4 , we write the nonlinear aircraft plant into the state form: Remark 6. The failure patterns in (30) include all “up to 1
failure” cases, but not all two-failure patterns are compensable
ẋ1 = (c1T 2
0 (x2 )x1 + 1 (x)) cos(x2 ) as shown in (31). The failure patterns in (30) and (31) implies
+ (c2T 2
+ that at least one piece of the elevator and one thrust should be
0 (x2 )x1 2 (x)) sin(x2 )
alive to guarantee the tracking objective. Since the system has
+ d1 g1 (x)u1 + d2 g1 (x)u2 + g31 (x)u3 + g41 (x)u4 , two pairs of control inputs which are similar physically, it is
  reasonable that those two actuators in one pair can compensate
1 each other. Moreover, after examining conditions (i), we also
ẋ2 = x4 − c1T 0 (x2 )x1 + 1 (x) sin(x2 )
x1 find out that the compensation between the two pairs is impos-
 
1 sible. In fact, if both of the elevator segments or both of the
+ c2 0 (x2 )x1 + 2 (x)
T
cos(x2 ) thrusts fail, the system actually loses one control freedom such
x1
that the tracking objective of two outputs cannot be ensured,
+ d1 g2 (x)u1 + d2 g2 (x)u2 + g32 (x)u3 + g42 (x)u4 , though there are two actuators remaining. Therefore, when de-
signing a compensation control for “up to two failure” cases
ẋ3 = x4 ,
for the aircraft model (28), naturally we divide the controls into
ẋ4 = T (x) + b1 x12 u1 + b2 x12 u2 , (28) two groups which consist of the physically similar actuators in
order to achieve two output tracking.
where
Next condition (ii) in Theorem 1 will be checked for those
0 (x2 ) = [x2 , x2 , 1] ,
2 T
failure patterns in (30) and (31) to see if they are compensable
1 (x) = p0 sin(x3 ), failure patterns.
2 (x) = p1 x4 x1 + p0 cos(x3 ),
2 It can be seen that the nonlinear actuation functions for u1 and
u2 , i.e., [d1 g1 (x), d1 g2 (x), 0, b1 x12 ]T and [d2 g1 (x), d2 g2 (x),
g1 (x) = a1 x12 cos(x2 ) + a2 x12 sin(x2 ),
0, b2 x12 ]T , are parallel (as b1 /b2 = d1 /d2 follows from (27)),
g2 (x) = −a1 x1 sin(x2 ) + a2 x1 cos(x2 ), which implies that the condition in Proposition 1 is satisfied so
g31 (x) = cos 1 cos(x2 ) + sin 1 sin(x2 ), that u1 and u2 will not enter the zero dynamics. In other words,
whatever the control inputs (v1 and v2 ) designed for those two
g41 (x) = cos 2 cos(x2 ) + sin 2 sin(x2 ), actuators and the possible failure signals (ū1 (t) and ū2 (t)) are,
sin(x2 ) cos(x2 ) the zero dynamics are fixed with respect to u1 and u2 .
g32 (x) = − cos 1 + sin 1 , On the other hand, for u3 and u4 , that condition is not satis-
x1 x1
sin(x2 ) cos(x2 ) fied such that the zero dynamics depend on the control design
g42 (x) = − cos 2 + sin 2 , which is based on a specific actuation scheme to handle the
x1 x1
redundancy. Here we choose equal control design for the two
(x) = [x12 x2 , x12 x22 , x12 , x12 x4 ]T (29) control inputs, that is, v3 = v4 , and next we show the resulting
1878 X. Tang et al. / Automatica 43 (2007) 1869 – 1883

zero dynamics are locally ISS with some requirements on the where v1 (t) = ū4 (t) cos 2 , v2 (t) = ū4 (t) sin 2 for case (i),
failure signals. v1 (t) = ū3 (t) cos 2 , v2 (t) = ū3 (t) sin 2 for case (ii), v1 (t) = 0,
Zero dynamics: With condition (24) satisfied for u1 and u2 , v2 (t) = 0 for case (iii).
the property of the zero dynamics (which actually is character- It will be seen that the zero dynamics are locally ISS with the
ized by the state x2 , because y1 = x3 , ẏ1 = x4 , and y2 = x1 ) is real aircraft constants and the parameters obtained in a certain
unchanged by the actuator failure resulting from u1 or u2 . After operation condition (Miller & William, 1999): S = 39.02 m2 ,
choosing the equal actuation scheme for u3 and u4 as v3 = v4 , c=1.98 m, and 1 =arctan 1216 53
, 2 =arctan 452
, =0.7377 kg/m3
we can specify the zero dynamics for three possible cases, that ◦
at the altitude of 5000 m, and for the 0 flap setting,
is, (i) u3 is alive but u4 fails, (ii) u4 is alive but u3 fails, (iii)
Cx1 = 0.39, Cx2 = 2.9099, Cx3 = −0.0758,
both u3 and u4 are alive.
However, it is noted from (28) that all the inputs appear in Cx4 = 0.0961,
the state equation of x2 , which means we have to consider the Cz1 = −7.0186, Cz2 = 4.1109, Cz3 = −0.3112,
controls and actuator failures in that state coordinates for study
of the stability of the zero dynamics. In order to avoid such a Cz4 = −0.2340, Cz5 = −0.1023,
difficult situation, we look for a change of coordinate  = T (x) Cm1 = −0.8789, Cm2 = −3.8520, Cm3 = −0.0108,
such that:
Cm4 = −1.8987, Cm5 = −0.6266. (36)
jT jT jT 2
g1 (x) + g2 (x) + kx = 0, Analyzing the unforced zero dynamics at x3 = 0, x4 = 0 and
jx1 jx2 jx4 1
x1 = Ve , which is the desired equilibrium point of the velocity,
jT that is, the system
(C cos(x2 ) + S sin(x2 ))
jx1   
  ˙ = SV e (c1T − q1 ¯ ) 0 arctan
T S
jT sin(x2 ) cos(x2 )
+ −C +S = 0, (32) C
jx2 x1 x1  

− arcsin √
where k = b1 /d1 = b2 /d2 , C = cos 1 , S = sin 1 for case (i), Ve S 2 + C 2
C = cos 2 , S = sin 2 for case (ii), C = cos 1 + cos 2 , and   
S
− CV e (c2T − q2 ¯ ) 0 arctan
T
S = sin 1 + sin 2 for case (iii), to transform the zero dynamics
into the form of ˙ = (, , , , ūj ), j = 3 or 4 without any C
 
input in it, where  = [x1 , x3 , x4 ]T . 
Solve the partial differential equations (32) and obtain − arcsin √ , (37)
Ve S 2 + C 2
 = S(x1 cos(x2 ) − q1 x4 ) − C(x1 sin(x2 ) − q2 x4 ), (33) where ¯ = [1 , 2 , 3 ]T , Ve = 60 m/s, we get the equilibrium
point of the unforced zero dynamics for three cases: (i) e =
where q1 = a1 /k and q2 = a2 /k.
0.00354; (ii) e = 0.00354; (iii) e = 0.00355. Definition of ISS
The Jacobian matrix of P (x) = [x1 , T (x), x3 , x4 ]T ,
and local ISS can be found in Khalil (1996). As indicated by
⎡ ⎤
1 0 0 0 Lemma 5.4 in Khalil (1996), from the uniformly asymptotic
jP (x) ⎢ ⎢ p21 (x2 ) p22 (x1 , x2 ) 0 −(Sq 1 + Cq 2 ) ⎥
⎥ stability of the unforced system at an equilibrium point, we can
=⎢ ⎥ (34) conclude the local ISS of the original system in some neigh-
jx ⎣ 0 0 1 0 ⎦
borhood of that equilibrium point. Investigating the stability of
0 0 0 1 the unforced zero dynamics at the equilibrium point, we see
is nonsingular when x1  = 0 and x2  = ±/2 + arctan S/C, that the zero dynamics are locally asymptotically stable for all
where p21 (x2 ) = S cos(x2 ) − C sin(x2 ) and p22 (x1 , x2 ) = three cases. It implies that the zero dynamics are locally ISS
−x1 (S sin(x2 ) + C cos(x2 )). Note that x1 > 0 because of the in a neighborhood of e with [x1 , x2 , x3 ]T in a neighborhood
physical meaning of x1 , the velocity of the aircraft. Consider of [60, 0, 0]T as the inputs when |ū3 (t)|, |ū4 (t)| < 10 000N . In
that when x2 = ±/2 + arctan S/C the thrust is orthogonal fact ū3 (t) and ū4 (t) are always nonnegative because their phys-
to the velocity which in turn implies that the velocity is un- ical meaning is thrust.
controllable at that point. Upon normal operation and design
conditions in reality that |x2 |>/2 and thrust misalignment, Remark 7. From (35), it is observed that actuator failures rep-
resented by v1 (t) and v2 (t) enter the zero dynamics. The signals
, is very small, we conclude that [, ]T = T (x) is a diffeo-
v1 (t) and v2 (t) are independent of any system states, and can be
morphism.
considered as external disturbances. They are not the key to the
Within the new coordinates, the zero dynamics can be ex-
stability of the zero dynamics because as external disturbances,
pressed as
as long as they are bounded signals, the ISS of the zero dynam-
˙ = S(c1T + ics guarantees the boundedness of their states, that is,  in this
1 (x) − x1 sin(x2 )x4
2
0 (x2 )x1
T
example. The system states x1 , x2 , and x3 characterize the zero
− q1  (x) + v1 (t)) − C(c2T 0 (x2 )x12 + 2 (x) dynamics via the transformation defined by (33), while x1 , x3 ,
+ x1 cos(x2 )x4 − q2 T (x) + v2 (t)), (35) and x4 are the inputs to the zero dynamics. Since the ISS of the
X. Tang et al. / Automatica 43 (2007) 1869 – 1883 1879

zero dynamics is a local property, the failure signals should be of the twin otter aircraft (28) can be developed to achieve output
restricted by some bounds (e.g., 0  ū3 (t)10 000N ) to ensure tracking of the pitch angle and velocity.
the states in the neighborhood of the equilibrium point. A common type of failure is an actuator sticking at a nonzero
value, resulting in an elevator segment or an engine thrust being
Hence all the sufficient conditions proposed in Theorem 1 stuck at a particular value. Therefore, we consider the constant
are satisfied, which means that the failure patterns in (30) and failure model, ūj (t) = ūj with ūj unknown, j = 1, 2, 3, 4.
(31) are compensable. For such a compensable failure pattern The adaptive controller is designed as
set, we can develop adaptive control schemes to compensate
actuator failures and guarantee the tracking control objective. ˆ ym1 , ẏm1 , ÿm1 )
(x, ,
Next, we will present an adaptive compensation controller for vi (t) = k̂1,i (t) + k̂2,i (t), i = 1, 2,
x12
the twin otter aircraft model.
Adaptive compensation design: With the structure of actua- 1
v3 = v4 = T
v0 (x, ,
ˆ ym2 , ẏm2 ), (38)
tors u1 and u2 satisfying condition (24) and the equal actua- (x2 )(t)
ˆ
tion scheme chosen for u3 and u4 , an adaptive compensation
scheme which can handle all the compensable failure patterns where k̂1 = [k̂1,1 , k̂1,2 ]T and k̂2 = [k̂2,1 , k̂2,2 ]T are the estimates
in (30) and (31) for the nonlinear longitudinal dynamic model of k1 = [k1,1 , k1,2 ]T and k2 = [k2,1 , k2,2 ]T which satisfy the

65
x1 (m/sec)

60

55
0 50 100 150 200 250 300 350 400 450 500
Velocity x1 (m/sec) vs time t (sec)

0.06
x2 (m/sec)

0.04

0.02

0
0 50 100 150 200 250 300 350 400 450 500
Attack angle x2 (rad) vs time t (sec)

0.1

0.05
x3 (rad)

−0.05

−0.1
0 50 100 150 200 250 300 350 400 450 500
Pitch angle x3 (rad) vs time t (sec)

0.02

0
x4 (rad/sec)

−0.02

−0.04

−0.06
0 50 100 150 200 250 300 350 400 450 500
Pitch rate x4 (rad/sec) vs time t (sec)

Fig. 1. System state variables.


1880 X. Tang et al. / Automatica 43 (2007) 1869 – 1883

matching conditions: where  is a known positive constant such that T


 > 0, and
k1T (I −  )[b1 , b2 ] = 1,
(1) T
(x) = [ T 2 T 2
0 (x2 )x1 cos(x2 ), 0 (x2 )x1 sin(x2 ),
k2T (I −  )[1, 1] = −diag{b1 , b2 } [ū1 , ū2 ]
(1) T (1) T
(39) x4 x12 sin(x2 ), x12 cos(x2 )v1 , x12 sin(x2 )v1 ,
for (1) and (2) being 2 × 2 diagonal matrices as the di- x12 cos(x2 )v2 , x12 sin(x2 )v2 , x12 cos(x2 ), x12 sin(x2 ),
agonal submatrices of a certain failure pattern matrix  =
diag{(1) , (2) }, ˆ = [ˆ 1 , ˆ 2 ]T is the estimate of  = [1 , 2 ]T cos(x2 ), sin(x2 )]T ∈ R 15 (43)
which satisfy the matching condition:
  with z2 z3 ,  and v0 given by
cos 1 cos 2
 = (I − (2) )[1, 1]T ,  = (40) z1 = x3 − ym1 ,
sin 1 sin 2
z2 = x4 + c1 z1 − ẏm1 ,
is the misalignment matrix, (x2 ) = [cos(x2 ), sin(x2 )]T ,
z3 = x1 − ym2 ,
and ˆ is the estimate of ,  ˆ is the estimate of  =
[c1T , c2T , p1 , d1 a1 , d1 a2 , d2 a1 , d2 a2 , T ] ∈ R 15 ,  = −c2 z2 − z1 − T ˆ − c1 x4 + c1 ẏm1 + ÿm1 ,
  v0 = − c3 z3 − p0 (cos(x3 ) sin(x2 ) − sin(x3 ) cos(x2 ))
d1 a1 d2 a1 0
 = d1 a2 d2 a2 0 ū. (41) − T

ˆ + ẏm2 , (44)
0 0 
1 > 0, 2 > 0, 3 > 0, 1i > 0, 2i > 0, i = 1, 2, and ci > 0,
Remark 8. Note that T (x2 ) = 0 if and only if x2 = ±/2 + i = 1, 2, 3.
arctan S/C. At that point, the velocity is uncontrollable. Under The adaptive compensation scheme applied to the aircraft
normal conditions as we discussed above, T (x2 )  = 0, and model (28), in the presence of actuator failures whose patterns
moreover, T (x2 ) > 0 due to small 1 and 2 and x2 which is belong to the compensable failure pattern set given in (30)
usually between −/4 and /4 according to its physical mean- and (31), ensures that closed-loop signals are bounded and the
ing. This property will be considered as a basis of employing output tracking error y(t) − ym (t) = [y1 − ym1 , y2 − ym2 ]T goes
projection in the design of the adaptive law for ˆ to avoid sin- to zero as t goes to infinity.
gularity. Simulation results: Applying the controller (38) with the
adaptive laws (42) to the nonlinear longitudinal model of the
ˆ ˆ and 
The adaptive laws for k̂1,i , k̂2,i , i = 1, 2, , ˆ are twin otter aircraft, we give the simulation results to illustrate
derived as the effectiveness of the compensation design proposed above.
k̂˙ = −sign[b ] z , i = 1, 2,
1i i 1i 2
For the simulation, the initial conditions are

k̂˙ 2i = −sign[bi ]2i z2 x12 , i = 1, 2, x(0) = [55, 0, 0.05, 0]T , ˆ


(0) = [0, 0, −0.01, 0]T ,
˙ˆ = proj[sign[ T ] T ˆ ,] (z3 2 v0 ), k̂1 (0) = [−1.2, −0.8]T , k̂2 (0) = [0, −0.05]T ,
˙ ˙ˆ = z3 3 (x),
ˆ = z2 1 (x),  (42) (0)
ˆ = [2, 0]T ,

0.1
y1 and y1m (rad)

0.05

−0.05

−0.1
0 50 100 150 200 250 300 350 400 450 500
Output y1 (solid) and reference y1m (dashed) (rad)

65
y2 and y2m (rad)

60

55
0 50 100 150 200 250 300 350 400 450 500
Output y2 (solid) and reference y2m (dashed) (rad)

Fig. 2. System outputs and reference outputs.


X. Tang et al. / Automatica 43 (2007) 1869 – 1883 1881

0.2

0.1

y1−y1m (rad)
0

−0.1

−0.2
0 50 100 150 200 250 300 350 400 450 500
Tracking error y1−y1m (rad)

5
y2−y2m (m/sec)

−5
0 50 100 150 200 250 300 350 400 450 500
Tracking error y2−y2m (m/sec)

0.2
u1 and u2 (rad)

0.1

−0.1

−0.2
0 50 100 150 200 250 300 350 400 450 500
Control inputs (elevator angle (rad)) u1(t) (solid), u2(t) (dashed)

4000
u3 and u4 (N)

2000

−2000
0 50 100 150 200 250 300 350 400 450 500
Control inputs (thrust (N)) u3(t) (solid), u4(t) (dashed)

Fig. 3. Tracking errors and control inputs.

(0)
ˆ = [0.5, 3, −0.1, −7, 4, −0.5, −0.1, The failure model considered in the simulation is that u1 fails
0.05, −0.15, 0.05, −0.1, 0, 0, 0, 0]T . (45) at 150 s with an unknown failure value 0.04 (rad) and u4 fails
at 300 s with an unknown failure value 1000 (N).
The gains in the adaptive laws are chosen as c1 = 1, c2 = 1, Fig. 1 shows the system response of all the state variables.
c3 =1, 1,i =1, 2,i =0.005, i =1, 2, 1 =0.001I , 2 =0.001I , The outputs and their corresponding reference are shown in
3 = 0.001I . Fig. 2. Fig. 3 shows the tracking errors of the two outputs and
The reference signals ym (t) = [ym1 (t), ym2 (t)]T to be the control signals of the four inputs. It can be seen from the
tracked are generated from two reference systems Wm1 (s) simulation results that the closed-loop system is stable in the
and Wm2 (s), respectively, that is, ym1 (t) = Wm1 (s)[r1 ](t), sense that all signals are bounded and that the asymptotic output
ym2 (t) = Wm2 (s)[r2 ](t), where Wm1 (s) = 1/(s 2 + 5s + 6) tracking is ensured even though there are two actuators failing
with r1 (t) = 0.1 sin(0.05t) as the reference input, and during operation whose failure time instants, failure values and
Wm2 (s)=1/(s +1) with r2 (t)=60+sin(0.05t) as the reference failure pattern are unknown to the adaptive failure compensa-
input. tion controller.
1882 X. Tang et al. / Automatica 43 (2007) 1869 – 1883

5. Conclusions Khalil, H. K. (1996). Nonlinear systems. (2nd ed.), Upper Saddle River, NJ:
Prentice-Hall.
The use of multiple actuating signals for control of dynamic Krstić, M., Kanellakopoulos, I., & Kokotović, P. V. (1995). Nonlinear and
adaptive control design. New York, NY: Wiley.
systems provides both the desirable control efficiency and actu- Marino, R., & Tomei, P. (1995). Nonlinear control design: Geometric,
ator fault tolerance capacity. Effective control of systems with adaptive, robust. Upper Saddle River, NJ: Prentice-Hall.
multiple actuating signals in the presence of uncertain failures Maybeck, P. S. (1999). Multiple model adaptive algorithms for detecting and
is important for many performance-critical systems such as air- compensating sensor and actuator/surface failures in aircraft flight control
craft, space structures, autonomous intelligent systems. There systems. International Journal of Robust and Nonlinear Control, 9(14),
1051–1070.
are key technical issues in the design of such control systems, Miller, R. H., & William, B. R. (1999). The effects of icing on the longitudinal
such as cooperation of multiple actuating signals and their fail- dynamics of an icing research aircraft. In Proceedings of the 37th AIAA
ure compensation. This paper has studied adaptive actuator fail- aerospace sciences meeting (pp. AIAA-99-0636), Reno, NV, January 1999.
ure compensation for a class of nonlinear multivariable sys- Mirkin, B. M., & Gutman, P. (2005). Model reference adaptive control of state
tems. Design conditions were derived based on which stable delayed system with actuator failures. International Journal of Control,
78(3), 186–195.
adaptive controllers were developed which are able to ensure Tang, X. D., Tao, G., & Joshi, S. M. (2002). Adaptive actuator failure
asymptotic tracking in the presence of uncertain failures. Such compensation for parametric strict feedback systems and an aircraft
adaptive control schemes and the associated design conditions application. Automatica, 39(11), 1975–1982.
are dependent on grouping of actuators and also on distribution Tao, G. (2003). Adaptive control design and analysis. New York, NY: Wiley.
of actuation among actuators. Those using fixed grouping and Tao, G., Chen, S. H., Tang, X. D., & Joshi, S. M. (2004). Adaptive control
of systems with actuator failures. New York, NY: Springer.
proportional actuation were investigated in this paper and were Vemuri, A. T., & Polycarpou, M. M. (1997). Robust nonlinear fault diagnosis
applied to control of a nonlinear longitudinal aircraft model. in input–output systems. International Journal of Control, 68(2), 343–360.
Stability and asymptotic tracking properties were proved in the- Wang, H., & Daley, S. (1996). Actuator fault diagnosis: An adaptive
ory, and desired system performance was verified by simulation observer-based technique. IEEE Transactions on Automatic Control, 41(7),
results. 1073–1078.
Wise, K., Brinker, J. S., Calise, A. J., Enns, D. F., & Elgersma, M. R. (1999).
Direct adaptive reconfigurable flight control for a tailless advanced fighter
References aircraft. International Journal of Robust and Nonlinear Control, 9(14),
999–1012.
Ahmed-Zaid, F., Ioannou, P., Gousman, K., & Rooney, R. (1991). Wu, N. E., Zhang, Y., & Zhou, K. (2000). Detection, estimation, and
Accommodation of failures in the f-16 aircraft using adaptive control. accommodation of loss of control effectiveness. International Journal of
IEEE Control Systems Magazine, 11(1), 73–78. Adaptive Control and Signal Processing, 17(7), 775–795.
Bodson, M., & Groszkiewicz, J. E. (1997). Multivariable adaptive algorithms Zhang, X., Parisini, T., & Polycarpou, M. M. (2004). Adaptive fault-tolerant
for reconfigurable flight control. IEEE Transactions on Control Systems control of nonlinear uncertain systems: An diagnostic information-based
Technology, 5(2), 217–229. approach. IEEE Transactions on Automatic Control, 49(8), 1259–1274.
Boskovic, J. D., & Mehra, R. K. (1999). Stable multiple model adaptive flight
control for accommodation of a large class of control effector failures.
In Proceedings of the 1999 American control conference (Vol. 6, pp. Xidong Tang received his B.S. and M.S. degree
1920–1924), San Diego, CA, June 1999. in Electrical Engineering from Shanghai Jiao
Tong University, China, in 1997 and 2000, re-
Boskovic, J. D., Yu, S.-H., & Mehra, R. K. (1998a). Stable adaptive fault-
spectively. He received his Ph.D. degree in Elec-
tolerant control of overactuated aircraft using multiple models, switching trical Engineering from University of Virginia
and tuning. In Proceedings of the 1998 AIAA guidance, navigation and in 2005. He is currently a senior research engi-
control conference (Vol. 1, pp. 739–749), Boston, MA, August 1998. neer at GM R&D Center at Warren, Michigan.
Boskovic, J. D., Yu, S.-H., & Mehra, R. K. (1998b). A stable scheme for His research interests include fault detection and
automatic control reconfiguration in the presence of actuator failures. In identification, fault-tolerance, vehicle diagnosis
Proceedings of the 1998 American control conference (pp. 2455–2459), and prognosis, system estimation, adaptive con-
Philadelphia, PA, June 1998. trol, nonlinear systems, multivariable control,
Calise, A. J., Lee, S., & Sharma, M. (2001). Development of a reconfigurable and control applications to vehicle systems.
flight control law for tailless aircraft. AIAA Journal of Guidance, Control,
and Dynamics, 25(5), 896–902. Gang Tao received his B.S. (EE) degree from
Chen, S. H., Tao, G., & Joshi, S. M. (2004). An adaptive actuator failure University of Science and Technology of China
compensation controller for MIMO systems. International Journal of in 1982, M.S. (EE, CpE and AM) degrees and
Ph.D. (EE) degree from University of Southern
Control, 77(15), 1307–1317.
California during 1984–1989. He worked in the
Corradini, M. L., & Orlando, G. (2003). A sliding mode controller for actuator areas of adaptive control, with particular inter-
failure compensation. In Proceedings of the 42nd IEEE conference on ests in adaptive control of systems with multiple
decision and control (pp. 4291–4296), Maui, Hawaii, December 2003. inputs and outputs and with nonsmooth non-
Demetriou, M. A., & Polycarpou, M. M. (1998). Incipient fault diagnosis linearities and actuator failures, in stability and
of dynamical systems using online approximators. IEEE Transactions on robustness of adaptive control systems, and in
Automatic Control, 43(11), 1612–1617. passivity characterizations of control systems.
Diao, Y., & Passino, K. M. (2001). Stable fault-tolerant adaptive fuzzy/neural He authored the 2003 Wiley textbook “Adaptive Control Design and Analysis”,
control for a turbine engine. IEEE Transactions on Control Systems co-authored the 1996 Wiley book “Adaptive Control of Systems with Actua-
tor and Sensor Nonlinearities” (with Petar V. Kokotovic), the 2003 Springer
Technology, 9(3), 494–509.
book “Control of Sandwich Nonlinear Systems” (with Avinash Taware) and
Ioannou, P. A., & Sun, J. (1996). Robust adaptive control. Upper Saddle the 2004 Springer book “Adaptive Control of Systems with Actuator Failures”
River, NJ: Prentice-Hall. (with Shuhao Chen, Xidong Tang and Suresh M. Joshi), and co-edited the
Isidori, A. (1995). Nonlinear control systems. (3rd ed.), New York, NY: 2001 Springer book “Adaptive Control of Nonsmooth Dynamic Systems”
Springer. (with Frank L. Lewis).
X. Tang et al. / Automatica 43 (2007) 1869 – 1883 1883

He is an associate editor for Automatica and a subject editor for International Dr. Joshi is a Fellow of the IEEE, the AIAA, and the ASME. His publications
Journal of Adaptive Control and Signal Processing (for which he was a guest include several articles and three books (in the areas of flexible spacecraft
editor for a 1997 special issue on adaptives systems with nonsmooth nonlin- control, and adaptive control in the presence of failures). He is the recipient
earities). He was an associate editor for IEEE Trans. on Automatic Control of: IEEE Control Systems Technology Award (1995), IEEE Judith A. Resnik
from 1996 to 1999. He organized and chaired the 2001 International Sympo- Award (2003), ASME-DSCD Charles S. Draper Award (2006), and IEEE
sium on Adaptive and Intelligent Systems and Control, held in Charlottesville, Region 3 Outstanding Engineer Award (2007). He also an amateur cartoonist
Virginia. He also organized invited sessions for 1996 IEEE CDC and 1999 and contributed the “Out of Control” cartoons to the IEEE Control Systems
IEEE CCA on adaptive control of systems with nonsmooth nonlinearities. Magazine from 1985 until 1994.

Suresh M. Joshi received is Ph.D. degree in


electrical engineering from Rensselaer Polytech-
nic Institute, Troy, NY, in 1973. He is presently
Senior Scientist for Control Theory at NASA-
Langley Research Center in Hampton, Virginia.
He also held visiting, adjunct, and research pro-
fessor positions at three universities. His re-
search interests include multivariable control
theory, nonlinear systems, adaptive control, and
applications to advanced aircraft and spacecraft.

You might also like