Robust Adaptive Control in The Presence of Unmodeled Actuator Dynamics

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Robust Adaptive Control in the Presence of Unmodeled

Actuator Dynamics

Jovan D. Bošković∗, Joseph A. Jackson† and Raman K. Mehra‡

In this paper we describe an adaptive controller design for a system where the open-loop plant dynamics
is unstable and the actuator dynamics is significant. The paper addresses the following related questions: (i)
How slow can the actuator dynamics become before they adversely affect the overall closed-loop stability?
(ii) When the baseline design is based on the assumption of fast actuator dynamics, what modifications to
the controller assure the overall system robustness in cases when this assumption is violated? The proposed
adaptive controller that addresses the item (ii) is of the Certainty Equivalence Adaptive Controller (CEAC)
type with Sequential Signal Filtering. We also show how this full-order adaptive controller can be combined
with a reduced-order controller designed for the case of fast actuator dynamics and moderate uncertainty.

I. Introduction

Adaptive control was established as a separate field of control systems theory in the late 1950s when first definitions
were given. While the interest in adaptive systems remained high in the 1960s and 1970s, it was not until 1980 that
adaptive control attracted wide interest when several proofs of stability were simultaneously published.1–3 Since then,
dozens of books have been published on adaptive control (early ones include4–7 ), along with hundreds of papers that
appeared in major journals. One of the key theoretical contributions was the discovery of a Lyapunov function that can
be used to design the adaptive control laws and prove the stability of the resulting adaptive control systems.4 This is
an asymptotic result that assures boundedness of the signals in the overall system, and asymptotic convergence of the
tracking error to zero under parametric uncertainty.
However, around the time when proofs of stability of adaptive control systems were given, the issue of robustness
of these controllers to plant model uncertainties and bounded disturbances was also raised. Several years later, a
paper by Rorhs et al.8 showed that the closed-loop adaptive control system can become unstable in the presence
of unmodeled dynamics. This prompted vigorous research aimed at assuring the robustness of the overall system to
bounded disturbances, time-varying parameters and some classes of unmodeled dynamics. The resulting schemes, such
as the dead zone, parameter projection, and σ-θ and e-θ modifications,4 were shown to assure robustness to bounded
disturbances and time-varying parameters. It also showed that such schemes assure robustness to unmodeled dynamics,
but only if the latter are small.7
Control design generally requires a sufficiently accurate model of the process or a plant to be controlled. The
resulting control design model is effectively a trade-off between the complexity of a high-fidelity process model, and
the simplicity required in order to arrive at analytically tractable control algorithms. Hence unmodeled dynamics
is invariably present in control design and implementation. One objective of the control designer is to arrive at a
sufficiently accurate model and control design that reduces the effects of unmodeled dynamics on the system in most
operating regimes of interest. While the design neglects unmodeled dynamics, the designer must still account for
their affects by making sure that the controller gains are sufficiently low to prevent excitation of high frequencies
of unmodeled dynamics. A good example is encountered in flight control where the control design neglects aircraft
∗ Principal Research Engineer & Intelligent & Autonomous Control Systems Group Leader, 500 W. Cummings Park, Suite 3000, Woburn, MA,

AIAA Senior Member, jovan@ssci.com


† Senior Research Engineer, 500 W. Cummings Park, Suite 3000, Woburn, MA, AIAA Senior Member, jjackson@ssci.com
‡ President & CEO, 500 W. Cummings Park, Suite 3000, Woburn, MA, AIAA Member, rkm@ssci.com

1 of 18

American Institute of Aeronautics and Astronautics


structural modes, but accounts for them by keeping the controller gains sufficiently low. Also, most flight control
designs neglect the effect of the actuators moving the flight control surfaces since their dynamics are fast.
While typical analysis and design tools for the case of adaptive control assumes that the unmodeled dynamics have
negligible effect on the system operation for most operational regimes, such an assumption may not be universally
justified. For instance, due to faults and failures (e.g. increased friction, electrical faults, or damage of flight control
actuators), actuator dynamics may become slow and their effect may become significant. In such cases the control
designs developed for the case of fast actuator dynamics may fail. Another challenge in flight control is to design
controllers that are robust to certain percentage of structural damage. As shown by the authors,14 certain cases of wing
damage may result in an unstable open-loop system, which makes the control design problem more difficult. When
the plant is unstable, the slow actuators can have a deleterious effect on the overall closed-loop system even when the
control design for the case without actuator dynamics stabilizes the system.
In this paper we consider the case when the open-loop plant dynamics is unstable and the controller is designed
to achieve the tracking objective for the case when actuator dynamics can be neglected. The paper addresses the
following related questions: (i) How slow can the actuator dynamics be before it adversely affects the overall closed-
loop stability?; and (ii) How should we modify the baseline design, based on the assumption of fast actuator dynamics,
to assure the overall system robustness in cases when this assumption is violated?
We describe in Section 2 a simple example where unmodeled actuator dynamics cause instability in a first order
plant. Specific stability bounds are derived and the analysis is illustrated through computer simulations. This is
followed by the design of a controller that assures the closed-loop stability in the case of significant actuator dynamics.
The related control strategy is referred to as the Certainty–Equivalence Adaptive Controller (CEAC) with Sequential
Signal Filtering. Section 4 proposes a strategy for switching between the low-order and full-order adaptive controllers
that assures closed-loop system stability.

II. Problem Statement

We consider the dynamics of a first order plant with uncertain constant parameters a and b, and actuator dynamics
with uncertain constant actuator gain ka of the form:

ẋ = ax + bua (1)
u̇a = −ka (ua − u), (2)

where x is the system state, ua denotes the state of the unmodeled actuator dynamics, and u is the control input vector.
Prior information includes the following: (a, b, ka ) ∈ Sp = {(a, b, ka ) : |a| ≤ ā, 0 < b ≤ b ≤ b̄, 0 < ka ≤ ka ≤ k̄a }
where ā, b, b̄, k a , and k̄a are known.
Reference Model: When actuator dynamics are significant, i.e. when they cannot be neglected, the closed-loop
system in general cannot follow a first-order reference model in the sense that the tracking error converges to zero
asymptotically from any initial condition. Hence we consider a second-order reference model of the form:

ẋm1 = xm2 (3)


ẋm2 = −k2 xm2 − k1 (xm1 − r), (4)

where r is a bounded piecewise continuous reference input such that |r(t)| ≤ r̄, ∀t ≥ 0.
Control Objective: The objective is to design a control signal u(t) such that limt→∞ [x(t) − xm1 (t)] = 0.
Comment: The above control problem is not trivial due to the fact that ua is not measurable and ka is uncertain. Hence,
even to stabilize the above plant, the feedback signal needs to account for the uncertainty in ka , as shown below.
Ideal Controller: We first consider the case when a, b and ka are known. Assuming zero initial conditions, we take a
derivative of (1) and use (2) and (1) to obtain:

ẍ + (ka − a)ẋ − ka ax = bka u. (5)

2 of 18

American Institute of Aeronautics and Astronautics


It is seen that, due to the term ka − a, simple feedback u = −kx cannot stabilize the closed-loop system globally,
regardless of the value of k. Hence, if the parametric uncertainty a is positive and large with respect to the actuator
speed (i.e. its time constant ka ), the closed loop system with the simple feedback law can become unstable. It can be
concluded that, to stabilize the system globally, the feedback control law needs to account for the term containing ẋ.
Hence the control law is modified as follows:
u = −kx − k0 ẋ.
The problem here is that ẋ depends on ua , and the latter is not measurable. Hence a simple observer for ua is built of
the form:
û˙ a = −ka (ûa − u),
where ûa denotes the estimate of ua . We recognize that, since ka > 0, the error ũa = ûa − ua tends to zero
exponentially.
The control law is now chosen as:
1
u= [−(ka a + k1 )x − (k2 + ka − a)(ax + ûa ) + k1 r],
bka
where the idea is to cancel the terms with x and ẋ, and add the reference model dynamics. Its use results in the
closed-loop system of the form:
ẍ + k2 ẋ + k1 x = k1 r − (k2 + ka − a)ũa . (6)
We next express the reference model in the form:
ẍm1 + k2 ẋm1 + k1 xm1 = k1 r,
and subtract from (6) to obtain:
ë + k2 ė + k1 e = −(k2 + ka − a)ũa , (7)
where e = x − xm1 , denoting the tracking error. Since the polynomial on the left-hand side of the above expression is
stable and limt→∞ ũa = 0, it follows that limt→∞ e(t) = 0 where the convergence to zero is exponential.
This simple analysis is possible due to the assumption that a, b and ka are known. However, when these parameters
are uncertain, the control problem becomes much more complex. The next section addresses how to approach the case
given uncertain parameters.

III. Case of Uncertain Parameters

In the case of uncertain a, b and ka the problem becomes difficult since we need to simultaneously estimate both
the state ua and the uncertain parameters.

A. Neglected Actuator Dynamics

We first simplify the problem by assuming that b is known (for this example, b = 1), and consider the case when the
actuator dynamics are fast, i.e. ka >> 1. In such a case, (2) can be written as:
1
u̇a = u − ua .
ka
Using the Singular Perturbation arguments, we can now conclude that ua ∼ = u, and the actuator dynamics are neglected.
Hence ẋ ∼= ax + u, and the objective reduces to finding a u(t) such that the state of the plant follows that of the first-
order reference model given by ẋm = −km (x − r), where km > 0. In such a case the adaptive control problem can be
solved using a standard approach.4 Since we assumed b = 1, the control law is chosen as:
u = θx + km r, (8)

3 of 18

American Institute of Aeronautics and Astronautics


where θ is an adjustable controller parameter, resulting in the following closed-loop system:

ẋ = (a + θ)x + km r.

The so-called matching condition is of the form: a + θ∗ = −km , where θ∗ is a desired value of θ, and the resulting
error equation is of the form:

ė = −km e + θ̃x, (9)

where e = x − xm denotes the tracking error, and θ̃ = θ − θ∗ denotes the parametric error. The adaptive law:
˙
θ̃ = θ̇ = −γθ ex (10)

can be shown to result in a stable system such that limt→∞ e(t) = 0. The stability can be shown4 using a Lyapunov
function:

1 2 θ̃2
V (e, θ̃) = (e + ).
2 γθ

We next note that, even when the assumption that the actuator dynamics are fast is justified, the above adaptive
control strategy can result in instability if the uncertainty in a is positive, i.e. when the open-loop plant dynamics are
unstable and a is large. We recall the expression for ẍ, (5). It is seen that, when a > ka , the control law (8) cannot
stabilize the closed loop system, even when ka >> 1 since it does not have a term in ẋ. This is true even if the modified
adaptive laws, designed to cope with external disturbances and some classes of unmodeled dynamics, are used. Such
adaptive laws include:
Adaptive Law with σ-θ Modification:9
˙
θ̃ = −γex − σθ, (11)

where σ > 0.
Adaptive Law with e-θ Modification:10
˙
θ̃ = −γex − ρ|e|θ, (12)

where ρ > 0.
Adaptive Law with Parameter Projection:7
˙
θ̃ = Proj[θ∈Sp] {−γex}, (13)

where Sp is defined previously.


Variable Structure Adaptation:11, 13

θ̄ex
θ=− , (14)
δ + |ex|

where δ > 0 and θ̄ = ā + km .


We next give an example to show that the above adaptive laws result in closed-loop system instability if a > ka .
Example 1: Let a = −1 initially, and a(t) = 2 for all t ≥ 2. Let also xm (0) = 1 and r(t) = 1 for all time, so
that xm (t) = 1 for all time. We next choose km = 1, γθ = 1, and ka = 1.9999, and simulate the system using the
controller (8) and adaptive law (13). The response of the system is given in Figure 1. Here we demonstrate that the
closed-loop system is unstable even though the adaptive law keeps the controller parameter bounded.

4 of 18

American Institute of Aeronautics and Astronautics


100
x(t)

x(t), xm (t)
50 xm (t)

−50
0 20 40 60 80 100 120 140 160 180 200

200

u(t) 0

−200
0 20 40 60 80 100 120 140 160 180 200

0 θ(t)
θ∗
θ(t), θ ∗

−2

−4
0 20 40 60 80 100 120 140 160 180 200
Time[sec]

Figure 1. Instability of the plant controlled by (8) where θ is adjusted using (13): Case when a = 2 and ka = 1.9999.

B. Case when actuator time constant is greater than the uncertainty bound

We next focus on the following problem: Let ka > ā and pose the following question: Is there a simple adaptive
controller that will stabilize the closed loop?
It is well know that the gradient adaptive law cannot in general stabilize the closed loop in the presence of unmod-
eled dynamics, and that different modifications are needed to assure the system robustness.4 We first show through a
simple analysis that the gradient adaptive law with an inappropriately chosen adaptation gain can result in closed-loop
instability. To see this, we simplify the problem by assuming that b = 1, xm (0) = 1 and r(t) = 1 for all time, so that
xm (t) = 1 for all time. Assuming zero initial conditions, we take a derivative of (1) and use (2), (1) and (8) to obtain:

ẍ + (ka − a)ẋ + ka km x = ka θ̃x + ka km r,

˙
where the adaptive law is of the form θ̃ = −γθ ex. By noting that x = xm + e, we further have:

ẍ + (ka − a)ẋ + ka km x = ka θ̃(e + xm ) + ka km r,

˙
and θ̃ = −γθ (e2 + (x − xm )xm ). We next neglect the products of errors, and let η1 = x, η2 = ẋ and η3 = θ̃ so that
the dynamics of the homogeneous part of the system is of the form:

η̇ = Aη

where  
0 1 0
A =  −ka km −(ka − a) −ka  .
 

γθ 0 0
since xm (t) ≡ 1 for all time. The resulting characteristic polynomial is of the form:

s3 + (ka − a)s2 + ka km s + ka γθ ,

5 of 18

American Institute of Aeronautics and Astronautics


resulting in the following stability condition:

γ < km (ka − a).

Example 2: Let, as in Example 1, a = −1 initially, and a(t) = 2 for all t ≥ 2. Let also km = 1, ka = 3.0095.
Adaptation gain is chosen as γθ = 1 and violates the above stability condition by a very small margin. The response
of the resulting system is shown in Figure 2. It is seen that the closed-loop system is unstable.

10
x(t)
x(t), xm (t) xm (t)
5

−5
0 20 40 60 80 100 120 140 160 180 200

100
u(t)

−100
0 20 40 60 80 100 120 140 160 180 200

20
θ(t)
0 θ∗
θ(t), θ ∗

−20
−40
−60
0 20 40 60 80 100 120 140 160 180 200
Time[sec]

Figure 2. Instability of the plant controlled by (8) where θ is adjusted using gradient adaptive law (10): Case when a = 2, ka = 3.0095 and
γ = 1.

In the case when the adaptive law is modified using the sigma-theta modification4 of the form θ̃˙ = −γθ ex − σθ,
the stability condition becomes:

(σ + ka − a)(σ(ka − a) + ka km ) > ka km σ + ka γθ ,

or (ka − a)σ 2 + (ka − a)2 σ + ka (km (ka − a) − γθ ) > 0. The latter condition is satisfied for any σ > 0 since ka > a
and γθ < km (ka − a). However, this is the condition derived from a characteristic polynomial of the linearized system,
and does not hold in general. In fact, the system with σ > 0 can still become unstable, as shown below.
Example 3: We choose the same parameters as in Example 2, with ka = 3.0069. Controller parameter is adjusted using
(11) with σ = 0.01. As seen in Figure 3, the closed-loop system is unstable due to the controller parameter becoming
unbounded.
Next question is if the adaptive laws that keep the controller parameters bounded can stabilize the system. In order
to illustrate this case, we slightly modify the controller as:

u = θx + kr, (15)

where both θ and k are adjusted as:

θ̇ = −γ1 ex (16)
k̇ = −γ2 er. (17)

6 of 18

American Institute of Aeronautics and Astronautics


10
x(t)

x(t), xm (t)
5 xm (t)

−5
0 20 40 60 80 100 120 140 160 180 200

100

u(t) 0

−100
0 20 40 60 80 100 120 140 160 180 200

20
θ(t)
0 θ∗
θ(t), θ ∗

−20
−40
−60
0 20 40 60 80 100 120 140 160 180 200
Time[sec]

Figure 3. Instability of the plant controlled by (8) where θ is adjusted using (11): Case when a = 2, ka = 3.0069, γ = 1 and σ = 0.01.

This controller is implemented for the case when, in the plant equation ẋ = ax + bua , both a and b are uncertain.
The simulation scenario is similar to the previous cases. The plant parameter a switches from a = −1 to a = 2 at
t = 2 seconds into the simulation. Both parameters are adjusted using the gain of one, i.e. γi = 1, i = 1, 2. Actuator
constant ka switches from ka = 4 to ka = 2.1. Hence ka > a. As seen in Figure 4, the closed-loop system is unstable

50
x(t), xm (t)

x(t)
0 xm (t)

−50
0 20 40 60 80 100 120 140 160 180 200
200
u(t)

−200
0 20 40 60 80 100 120 140 160 180 200
0
θ(t)
θ(t), θ ∗

θ∗
−5

−10
0 20 40 60 80 100 120 140 160 180 200
3
k(t)
k(t), k∗

2 k∗
1
0
0 20 40 60 80 100 120 140 160 180 200

Figure 4. Instability of the plant controlled by (15) where θ(t) and k(t) are adjusted using (16) and (17): Case when a = 2, ka = 2.1,
γ1 = 1 and γ2 = 1.

7 of 18

American Institute of Aeronautics and Astronautics


even though both controller parameters are bounded.
We can conclude that, even when controller parameters are bounded and the actuator is faster than an unstable
plant, we can find cases when the closed-loop system will become unstable. Hence a different approach is needed in
the case of significant actuator dynamics and unstable plant dynamics.

C. Certainty Equivalence Adaptive Control with Sequential Signal Filtering

In order to solve this problem, we note that the a control strategy is needed that simultaneously estimates the non-
measurable state ua and the uncertain parameters a, b and ka . The main challenge here is that, in the case of slow
and uncertain actuator dynamics, the problem becomes that of adaptive control design for the case of unmatched
uncertainty, i.e. the case when the uncertainty and control input are separated by one or more integrators. We have
studied this problem in detail19 for the case of measurable states and developed a systematic adaptive control design
procedure. Here we include its extension to the case of non-measurable states.
We first introduce the following filtered variables:
1
xF = ·x
s + λF
1
uF = ·u
s + λF
1
uaF = · ua ,
s + λF
where λF > 0.
We now filter the equations (1), (2) to obtain:

x = (λF + a)xF + buaF (18)


ua = (λF − ka )uaF + ka uF , (19)

modulo exponentially decaying initial conditions. We next solve (18) for uaF , substitute the result into (19), and
substitute the resulting expression for ua into (1) to obtain:

ẋ = θ1 (x − λF xF ) + θ2 xF + θ3 uF , (20)

where
θ1 = λF + a − ka , θ2 = ka a, θ3 = bka .
Hence ua is eliminated from further analysis, and ẋ is expressed in terms of measurable or obtainable signals.
The expression (20) can also be rewritten as:

ẋ = θT ω, (21)

where θ = [θ1 , θ2 , θ3 ]T and ω = [x − λF xF , xF , uF ]T . Based on definition of parameters θi , we note that θ ∈ Sθ =


{θ : ci ≤ θi ≤ c̄i }, where c3 > 0.
Ideal Controller: We next differentiate (20) to obtain:

ẍ = θ1 (θT ω − λF (x − λF xF )) + θ2 (x − λF xF ) + θ3 (u − λF uF )
= θ1 θT ω − (θ1 λF − θ2 )(x − λF xF ) + θ3 (u − λF uF ).

The ideal controller is now chosen as:


1
u = λF uF + [−(θ1 + k2 )θT ω + (θ1 λF − θ2 )(x − λF xF ) − k1 (x − r)], (22)
θ3
which is expressed in terms of measurable or obtainable signals.

8 of 18

American Institute of Aeronautics and Astronautics


Keeping in mind that θT ω ≡ ẋ, it can now be readily shown that this controller results in limt→∞ e(t) = 0.
We now study the case of uncertain parameters, and build an observer to estimate the elements of ω on-line.
Observer: The observer is chosen in the form:

x̂˙ = θ̂T ω − λê, (23)

where λ > 0, and ê = x̂ − x denotes the estimation error.


Error Model: Upon subtracting (21) from the above expression we obtain:

ê˙ = −λê + θ̃T ω, (24)

where θ̃ = θ̂ − θ denotes the parameter error.


Adaptive Law: The adaptive law is now chosen in the form:
˙ ˙ Γωê
θ̂ = θ̃ = Proj[θ∈Sθ ] {− }, (25)
1 + ωT ω
where Γ = ΓT > 0. and Proj(·) {(·)} denotes the projection operator (see related reference18). An important property
of this operator is that it assures that θ̂(0) ∈ Sθ implies that θ̂(t) ∈ Sθ , ∀t ≥ 0. Another important property that will
be used in the proof of stability is that it assures that

(θ̂ − θ)T Γ−1 θ̃˙ + êθ̃T ω ≤ 0, ∀(ê, ω) and ∀(θ̂, θ) ∈ Sθ . (26)

Certainty–Equivalence Control Law: The adaptive control law is now chosen in the form:
1
u = λF uF + [−(θ̂1 + k2 )θ̂T ω + (θ̂1 λF − θ̂2 )(x − λF xF ) − k1 (x − r)]. (27)
θ̂3

Before proceeding to the proof of stability, we first note that the control law (27) can be rewritten as:
1
u = −(θ̂1 + k2 − λF )uF − (θ̂1 + k2 )θ̂2 xF
θ̂3
  
1
+ θ̂1 (θ̂1 + k2 ) − θ̂1 λF + θ̂2 ℓ(x) + k1 r , (28)
θ̂3
where ℓ(x) = x − λF xF . We now consider the following Proposition:
Proposition 2:
ω T ω̇
≤ cω , ∀ω ∈ IR3 , (29)
1 + ωT ω
where cω can be readily computed.
Proof: We recall that ℓ(x) = x − λF xF . Given that ω = [ℓ(x), xF , uF ]T , we further have:

ℓ(x)ℓ̇(x) ℓ(x)(ẋ − λF ℓ(x))


=
1 + ωT ω 1 + ωT ω
ℓ(x)((θ1 − λF )ℓ(x) + θ2 xF + θ3 uF )
=
1 + ωT ω
2
(a − ka )ℓ(x) + (θ2 xF + θ3 uF )ℓ(x)
= .
1 + ωT ω
We also have that
xF ẋF xF ℓ(x)
= ≤ 1.
1 + ωT ω 1 + ωT ω

9 of 18

American Institute of Aeronautics and Astronautics


We next use (28) to obtain:

uF u̇F uF (u − λF uF )
=
1 + ωT ω 1 + ωT ω

−(θ̂1 + k2 )θ̂3 u2F + (θ̂1 λF − θ̂2 θ̂12 − k2 θ̂1 − k1 )uF ℓ(x) − ((θ̂1 + k2 )θ̂2 + k1 λF )xF uF + k1 uF r
= .
θ̂3 (1 + ω T ω)

We now define:

(a − ka )θ̂3 ℓ(x)2 − (θ̂1 + k2 )θ̂3 u2F + (θ2 + 1)θ̂3 ℓ(x)xF


ζ(ω) =
θ̂3 (1 + ω T ω)
(θ̂1 λF − θ̂2 θ̂12 − k2 θ̂1 − k1 + θ3 θ̂3 )uF ℓ(x) − ((θ̂1 + k2 )θ̂2 + k1 λF )xF uF + k1 uF r
+ .
θ̂3 (1 + ω T ω)

Since θ ∈ Sθ , for a given range of x and u, the upper bound on cω can be computed numerically. 
Hence there exists a cω > 0 such that

ζ(ω) ≤ cω , ∀ω and ∀(θ, θ̂) ∈ Sθ . (30)

Theorem 1: If the observer gain λ is chosen to satisfy:


1
λ> cω , (31)
2
where cω is given by (30), then the following properties of the observer (23) and adaptive law (25) hold:

(a) √ ∈ L2 .
1 + ωT ω
˙
(c) θ̂ = θ̃˙ ∈ L2 ∩ L∞ , and
T
(d) θ̃˙ ω = α(ω)ê, where α(ω) is bounded for all t ≥ to .
Proof of Theorem 1: We first choose a tentative Lyapunov function of the form:

1 ê2
V (ê, θ̃) = [ + θ̃T Γ−1 θ̃].
2 1 + ωT ω
Its derivative along the solutions of (23), (25) yields:

λê2 + êθ̃T ω 1 ω T ω̇ ˙
V̇ (ê, θ̃) = − − + θ̃T Γ−1 θ̃
1 + ωT ω 2 (1 + ω T ω)2
ê2 1
≤ − T
[λ − cω ]
1+ω ω 2
≤ 0, ∀(ê, θ̃) 6= (0, 0),

where the property (26) was used. It follows that ê/ 1 + ω T ω and θ̃ are bounded. Integrating V̇ from 0 to ∞ we
obtain:
Z ∞
ê(t)dt
V (∞) − V (0) = −λ .
0 1 + ω T (t)ω(t)

Since both V (∞) and V (0) are bounded, it follows that ê/ 1 + ω T ω ∈ L2 .

10 of 18

American Institute of Aeronautics and Astronautics


˙ ˙ Γωê
We next note that, from the adaptive law (25), we have that θ̃ is either zero, or θ̃ = − 1+ω T ω , which can be rewritten

as:
˙ Γω ê
θ̃ = − √ ·√ .
1 + ωT ω 1 + ωT ω

Hence it follows that θ̃˙ ∈ L∞ ∩ L2 .


We next consider two definitions relevant for the proof of stability. Let PC [0,∞) be a set of all real piecewise
continuous functions defined on the interval [0, ∞) which have bounded discontinuities. Theorem 1 will be proven
using the growth rates of unbounded signals which belong to PC [0,∞) . We next consider the following definitions:4
Definition 1: Let (x, y) ∈ PC [0,∞) .
(a) If there exist positive constants c1 , c2 and to ∈ IR+ such that |y(t)| ≤ c1 |x(t)| + c2 , ∀t ≥ to , then y(t) is large
order x(t), denoted by y(t) = O[x(t)].
(b) If there exist a function β(t) ∈ PC [0,∞) and to ∈ IR+ such that |y(t)| = β(t)|x(t)| and limt→∞ β(t) = 0, then
y(t) is small order x(t), denoted by y(t) = o[x(t)].
(c) If y(t) = O[x(t)] and x(t) = O[y(t)] then x and y are said to be equivalent, i.e. x ∼ y, and grow at the same rate.
Definition 2: Let the system be described by ẋ = A(t)x + B(t)r where r(t) and the elements of A(t) and B(t) are
uniformly bounded. Then:

d
kxk ≤ kẋk ≤ k1 kxk + k2 ,
dt

where k1 , k2 ∈ IR+ , and kx(t)k can grow at most exponentially, i.e. there exist constants k3 , k4 , k5 ∈ IR+ such that

kx(t)k ≤ k3 exp(k4 t) + k5 .

Comment:
√ Since the adaptive law with normalization is used, we cannot prove that ê is bounded, but only that
ê/ 1 + ω T ω is bounded. Hence in the next theorem, growth rates of the signals in the closed-loop system need to be
used to prove the overall stability.
Theorem 2: Certainty–Equivalence Control Law (27) for the plant (1) with uncertain actuator dynamics (2), along with
the adaptive law (25) and observer (23), assures that all the signals in the system are bounded and that limt→∞ [x(t) −
xm1 (t)] = 0.
Proof of Theorem 2: We first introduce the following coordinate transformation:

η1 = x̂ (32)
T
η2 = θ̂ ω + θ̂1 ê. (33)

Upon recalling the definition of ℓ(x) = x − λF xF and taking derivatives of η1 and η2 we obtain:

η̇1 = η2 − λê
˙T ˙
η̇2 = θ̂ ω + θ̂1 (θT ω − λF ℓ(x)) + θ̂2 ℓ(x) + θ̂3 (u − λF uF ) + θ̂1 ê − λθ̂1 ê + θ̂1 θ̃T ω
˙T ˙
= θ̂ ω + θ̂1 θ̂T ω − (θ̂1 λF − θ̂2 )ℓ(x) + θ̂3 (u − λF uF ) + θ̂1 ê − λθ̂1 ê.

Substituting the control law (27) into the above equation results in:

˙T ˙
η̇2 = θ̂ ω − k2 θ̂T ω + θ̂1 ê − λθ̂1 ê − k1 (x − r).

Since η2 = θ̂T ω + θ̂1 ê and ê = x̂ − x, we further have:

˙T ˙
η̇2 = −k2 η2 − k1 (η1 − r) + θ̂ ω + ((k2 − λ)θ̂1 + θ̂1 + k1 )ê.

11 of 18

American Institute of Aeronautics and Astronautics


We define the errors η̃1 = η1 − xm1 η̃2 = η2 − xm2 and subtract the reference model to obtain:
" # " #
0 1 0
η̃˙ = η̃ + ˙ T ˙ .
−k1 −k2 θ̂ ω + ((k2 − λ)θ̂1 + θ̂ 1 + k1 )ê

˙
Since uncertain parameters are constant, θ̃˙ bounded implies that θ̂ is also bounded. Also, θ̃ bounded implies that θ̂ is
˙ Γωê
bounded as well. Since adaptive algorithms with projection are used, the term θ̂ is either zero, or is equal to − 1+ω Tω.

Hence the term θ̇ω can be expressed as ξo (ω)ê, where ξo (ω) is a bounded function of ω. Hence
" # " #
0 1 0
η̃˙ = η̃ + ê,
−k1 −k2 ξ

˙
where ξ = ξo + (k2 − λ)θ̂1 + θ̂1 + k1 is a bounded parameter. It is seen that the stability of the system can be proved
if we can prove that ê is bounded. The will be the main objective of the remainder of the proof.
We first note that the control law (28) is a linear function of ω and r, i.e. u = pTu ω + puo r, where p =
   ⊤
1 k1
θ̂1 (θ̂1 + k2 ) − θ̂1 λF + θ̂2 , −(θ̂1 + k2 ), −(θ̂1 + k2 − λF )θ̂3 and puo = .
θ̂3 θ̂3
We next take a derivative of ω to obtain:
   
θT ω − λF ω1 0
ω̇ =  ω1  +  0  r,
   

pTu ω − λF ω3 puo

which can be rewritten as:

ω̇ = Aω ω + bω r. (34)

Since θ̃ is bounded, the above system, along with:

ê˙ = −λê + θ̃T ω,

is a linear time varying system with bounded parameters in which the signals can grow at most exponentially.
√ √
We next note that, since ê/ 1 + ω T ω ∈ L2 , it follows that ê = β(t) 1 + ω T ω, where β ∈ L2 .
1
Let ω grow in an unbounded fashion. Since ê = ê(0) exp (−λt) + s+λ F
θ̃T ω and θ̃ is bounded, it follows that
ê = O(supτ ≤t kω(τ )k). Since ê˙ = −λê + θ̃ ω, it also follows that ê˙ = O(supτ ≤t kω(τ )k).
T

From (34) it also follows that ω̇ = O(supτ ≤t kω(τ )k), and ω = O(supτ ≤t kω(τ )k). Similarly, η = O(supτ ≤t kω(τ )k).
Hence all the signals grow at the same rate, i.e. supτ ≤t |ê(τ )| ∼ supτ ≤t |η(τ )| ∼ supτ ≤t kω(τ )k.

On the other hand, since ê = β(t) 1 + ω T ω, where β ∈ L2 , and ê˙ = O(supτ ≤t kω(τ )k), it follows that ê =
O(supτ ≤t kω(τ )k),4 i.e. ê and kωk grow at different rates, which is a contradiction. Hence all the signals in the
closed-loop system are bounded.
Since ω is bounded and 1+ωê T ω ∈ L2 , it follows that ê ∈ L2 . Also, ê˙ is bounded, which implies that limt→∞ ê(t) =
4
0. The latter conclusion also implies that limt→∞ η̃(t) = 0, which completes the proof. 
Example 4: We next consider the case when, initially, a = −1, b = 1 and ka = 4. All initial conditions are zero, and
r(t) = 1 for all time. The uncertainty set is Sp = {(a, b, ka ) = −1 ≤ a ≤ 1, 0.5 ≤ b ≤ 1, 1 ≤ ka ≤ 4}.

Reduced-order Adaptive Controller: The plant is initially controlled by a controller u = θx + kr, where θ̇ =
−γ1 (x − xm )x, and k̇ = −γ2 (x − xm )r, where xm is the output of the reference model ẋm = −xm + r. The gains are
chosen as γ1 = γ2 = 1, and parameter projection is used to keep the parameters within their bounds computed using

12 of 18

American Institute of Aeronautics and Astronautics


2

x(t), xm (t)
x(t)
1 xm (t)

0
0 5 10 15 20 25 30 35 40 45 50
2

u(t)
1

0
0 5 10 15 20 25 30 35 40 45 50

0 θ(t)
θ(t), θ ∗

−2 θ∗

−4
0 5 10 15 20 25 30 35 40 45 50
3
k(t)
k(t), k∗

2 k∗
1
0
0 5 10 15 20 25 30 35 40 45 50

Figure 5. Response of the plan controlled by the reduced-order adaptive controller: Nominal Case.

the uncertainty set bounds and matching conditions a + bθ∗ = −km and bk ∗ = km . The resulting system response is
shown in Figure 5. It is seen that the response is excellent and the control objective is met.
At t = 2 seconds the plant and actuator parameters switch to (a, b, ka ) = (1, 0.5, 1). Such a change in a makes
the open-loop plant unstable and the reduced-order controller cannot stabilize the plant due to slow actuator dynamics
and large parametric uncertainty in a. The response of the closed-loop system is shown in Figure 6. It is seen that the
system becomes unstable even though the controller parameters are bounded.
SSF-CEAC Simulation: In this case the objective is to follow a second-order reference model. The reference model
parameters are k1 = 4 and k2 = 2.8; the filter gain is λF = 10, observer gain is λ = 6 and adaptive gain is
Γ = 1000 · I. We note that such a large adaptive gain is possible since the adaptive laws include the normalization
term. The response of the system with the SSF-CEAC controller is shown in Figure 7. It is seen that, after an initial
transient, the closed-loop is stabilized and the performance is good.

13 of 18

American Institute of Aeronautics and Astronautics


10

x(t), xm (t)
x(t)
0 xm (t)

−10
0 5 10 15 20 25 30 35 40 45 50
50

u(t)
0

−50
0 5 10 15 20 25 30 35 40 45 50

0 θ(t)
θ(t), θ ∗

−2 θ∗

−4
0 5 10 15 20 25 30 35 40 45 50
3
k(t)
k(t), k∗

2 k∗
1
0
0 5 10 15 20 25 30 35 40 45 50

Figure 6. Response of the plan controlled by the reduced-order adaptive controller: Case when parameters switch to (a, b, ka ) = (1, 0.5, 1)
at t = 2 seconds.

4
x(t), xm1 (t)

x(t)
2 xm1 (t)

0
0 5 10 15 20 25 30 35 40 45 50
20
u(t)

−20
0 5 10 15 20 25 30 35 40 45 50
5
θ̂ 1(t), θ 1

θ̂ 1(t)
0 θ1
−5
0 5 10 15 20 25 30 35 40 45 50
5
θ̂ 2(t), θ 2

θ̂ 2(t)
0 θ2
−5
0 5 10 15 20 25 30 35 40 45 50
10
θ̂ 3(t), θ 3

θ̂ 3(t)
5 θ3
0
0 5 10 15 20 25 30 35 40 45 50

Figure 7. Response of the plan controlled by the full-order adaptive controller: Case when parameters switch to (a, b, ka ) = (1, 0.5, 1) at
t = 2 seconds.

14 of 18

American Institute of Aeronautics and Astronautics


IV. Switching Controller

In this section we propose a technique for combining the reduced-order and full-order controllers. The technique
is based on the Multiple Models, Switching, and Tuning (MMST) methodology illustrated in Figure 8.
The concept of MMST is based on the idea of describing the dynamics of the system using different models for dif-
ferent operating regimes; such models identify in some sense the current dynamics of the system and are consequently
referred to as the identification models. The basic idea is to set up such identification models and corresponding con-
trollers in parallel, Figure 8, and to devise a suitable strategy for switching among the controllers to achieve the desired
control objective. While the plant is being controlled using one of these controllers, the identification models are run in
parallel to generate some measure of the corresponding identification errors and find a model which is, in some sense,
closest to the current operating regime of the plant. Once such a model is found, the switching mechanism switches to
(or stays at) the corresponding controller, where the switching interval is a parameter chosen by the designer. The main

ON
ADAPTIVE

...
ESTIMATOR
BANK
... O1

SWITCHING BLOCK Plant

... C1
CONTROLLER
...

BANK
CN

Figure 8. Structure of the Multiple Model-Based Controller: Outputs of the parallel observers O1 , O2 , ... ON are used to find that closest
in some sense to the current plant dynamics, and switch to the corresponding controller.

feature of this approach is that in LTI systems, it results in a stable overall system in which asymptotic convergence of
the output error to zero is guaranteed under relatively mild conditions.12
The switching algorithm is implemented by first calculating the following performance indices:
Z t
Ij (t) = c1 kêj (t)k2 + c2 exp (−λo (τ − to ))kêj (τ )k2 dτ, j = 1, 2, ..., N, (35)
to

where êj = x̂j − x, c1 > 0, c2 > 0, and λo > 0. The scheme is now implemented by calculating and comparing the
above indices every ts instants, finding the minimum, and switching to the corresponding controller.
To implement this scheme, we first design an observer for the reduced-order case:

x̂˙ R = âx + b̂u − λR êR ,

where êR = x̂R − x. The observer for the full-order case is given by (23). Now the above indices are computed and
compared, and the scheme switches to the controller corresponding to the minimum of the performance index. The
properties of the resulting system are illustrated through the following example.

15 of 18

American Institute of Aeronautics and Astronautics


Example 6: We simulate the MMST-based scheme using the same plant and scenario as in Example 5. The scheme
starts with the reduced-order controller. As the plant and actuator parameters switch to the new values, the scheme
switches to the full-order controller. The overall response is shown in Figure 9. The switching sequence is shown

x(t), xm1 (t)


x(t)
2 xm1 (t)

0
0 5 10 15 20 25 30 35 40 45 50
20

u(t) 0

−20
0 5 10 15 20 25 30 35 40 45 50
5
θ̂ 1(t), θ 1

θ̂ 1(t)
0 θ1
−5
0 5 10 15 20 25 30 35 40 45 50
5
θ̂ 2(t), θ 2

θ̂ 2(t)
0 θ2
−5
0 5 10 15 20 25 30 35 40 45 50
10
θ̂ 3(t), θ 3

θ̂ 3(t)
5 θ3
0
0 5 10 15 20 25 30 35 40 45 50

Figure 9. Response of the plan controlled by the multiple model adaptive controller: Case when parameters switch to (a, b, ka ) = (1, 0.5, 1)
at t = 2 seconds.

in Figure 10, where “1” corresponds to the reduce-order controller, and “2” to the full-order one. It is seen that the
scheme switches to the full-order controller at t = 2.25 seconds. Hence the scheme does not unnecessarily use the full-
order controller, and achieves the control objective if there are large changes in the system that make the reduced-order
controller non-robust.

V. Conclusions

In the paper we consider the case when the open-loop plant dynamics is unstable and an adaptive controller is
designed to achieve the tracking objective in the case when the actuator dynamics can be neglected. The paper addresses
the following questions related to this case: (i) How slow does the actuator dynamics need to get before it will adversely
affect the overall closed-loop stability? and (ii) How should we modify the baseline design, based on the assumption
of fast actuator dynamics, to assure the overall system robustness in cases when this assumption is violated?
In the paper we carry out a thorough analysis of related instability mechanisms due to unmodeled actuator dynamics
on a simple example. Specific stability bounds are derived and the analysis is illustrated through computer simulations.
This is followed by the design that assures the closed-loop stability in the case of significant actuator dynamics, and a
strategy for switching between the low-order and high-order adaptive controllers is given.
The proposed full-order adaptive controller is of the Certainty Equivalence Adaptive Controller (CEAC) type with
Sequential Signal Filtering. The idea behind the CEAC is simple since it consists of the following two steps: (a) Design
a controller for the case of known parameters; and (b) Replace the parameters with their estimates at every instant.
However, there are challenges related to this approach. For example, what strategy should be devised for estimating
plant parameters on-line when the actuator output is not measurable, and how do we prove the overall system stability?
In the paper we have shown that a single first-order observer is sufficient to estimate the plant parameters even though
the uncertainty appears in both equations. This was brought about by proper signal filtering and by bringing all the
parameters in the plant dynamics equation. To estimate the parameters, adaptive algorithms with normalization and

16 of 18

American Institute of Aeronautics and Astronautics


3

2.5

Switching Sequence
1.5

0.5

0
0 5 10 15 20 25 30 35 40 45 50
Time [sec]

Figure 10. Response of the switching sequence. The fault occurs at 2 seconds, and the estimator switches at 2.25 seconds.

projection were used. This made the proof of stability non-trivial since the observer properties allowed us to prove that
only a normalized estimation error is bounded, rather than the error itself. For this reason, the second half of the proof
relied on the growth rates of the signals in the closed-loop system.
In the paper we have also shown how this full-order adaptive controller can be combined with a reduced-order con-
troller designed for the case of fast actuator dynamics and moderate uncertainty. Properties of the proposed approach
are illustrated through computer simulations.

References
1 K. S. Narendra, Y.-H. Lin and L. S. Valavani, ”Stable Adaptive Controller Design II - Proof of Stability”, IEEE Transactions on Automatic

Control, Vol. AC-25, pp. 440-448, June 1980.


2 A. S. Morse, ” Global Stability of Parameter Adaptive Control Systems”, IEEE Transactions on Automatic Control, Vol. AC-25, pp. 433-439,

June 1980.
3 G. Goodwin, P. Ramadge and P. Caines, ”Discrete-time Multivariable Adaptive Control”, IEEE Transactions on Automatic Control, Vol.

AC-25, pp. 449-456, June 1980.


4 K. S. Narendra and A. M. Annaswamy. Stable Adaptive Systems. Prentice-Hall Inc., Englewood Cliffs, NJ, 1989.
5 G. C. Goodwin and K. S. Sin, Adaptive Filtering Prediction and Control, Prentice Hall, Englewood Cliffs, NJ, 1984.
6 S. Sastry and M. Bodson, Adaptive Control: Stability, Convergence and Robustness, Prentice-Hall, Englewood Cliffs, NJ,1989.
7 P. Ioannou and K. Sun, Robust Adaptive Control, Prentice Hall Inc., Englewood Cliffs, NJ, 1996.
8 C. E. Rohrs, L. Valavani, M. Athans, and G. Stein, “Robustness of Continuous-time Adaptive Control Algorithms in the Presence of

Unmodeled Dynamics,” IEEE Transactions on Automatic Control, Vol. 30, no. 9, pp. 881-889, 1985.
9 P. A. Ioannou and P. V. Kokotović, “Instability Analysis and Improvement of Robustness of Adaptive Control,” Automatica, Vol. 20, no. 5,

pp. 583-594, 1984.


10 K. S. Narendra and A. M. Annaswamy, “Robust Adaptive Control in the Presence of Bounded Disturbances,” IEEE Transactions on Auto-

matic Control, Vol. 31, no. 4, pp. 306-315, 1986.


11 L. Hsu and R. R. Costa, ”Variable Structure Model Reference Adaptive Control using only Input and Output Measurements”, International

Journal of Control, Vol. 49, No. 2, pp. 399-416, 1989.


12 K. S. Narendra and J. Balakrishnan, “Adaptive Control using Multiple Models”, IEEE Transactions on Automatic Control, Vol. 42, No. 2,

pp. 171-187, February 1997.


13 K. S. Narendra and J. D. Bošković, “A Combined Direct, Indirect and Variable Structure Method for Robust Adaptive Control”, IEEE

Transaction on Automatic Control, Vol. 37, pp. 262-268, February 1992.

17 of 18

American Institute of Aeronautics and Astronautics


14 J. D. Bošković, N. Knoebel and J. Redding, “Integrated Damage Adaptive Control System (IDACS)”, Final Report under NASA Langley

Phase II SBIR, Contract No. NNL07AA02C, May 2009.


15 J. D. Bošković and N. Knoebel, “A Comparison Study of Several Adaptive Control Strategies for Resilient Flight Control”, (invited paper)

in Proceedings of the 2009 AIAA Guidance, Navigation & Control Conference, Chicago, IL, 10 - 13 August 2009.
16 J. D. Bošković, N. Knoebel and R. K. Mehra “An Initial Study of a Combined Robust and Adaptive Control Approach to Guaranteed

Performance Flight Control”, in Proceedings of the 2008 American Control Conference, Seattle, WA, June 11–13, 2008.
17 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.
18 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.
19 J. D. Bošković and Z. Han, “Certainty Equivalence Adaptive Control of Plants with Unmatched Uncertainty using State Feedback”, IEEE

Transactions on Automatic Control, Vol. 54, No. 8, pp. 1918-1924, August 2009.
20 J. D. Bošković, “Stable Certainty Equivalence Adaptive Control using Normalized Parameter Adjustment Laws”, in Proceedings of the

2009 American Control Conference, St. Louis, MO, June 10-12, 2009.
21 J. D. Bošković and N. Knoebel, “Adaptive Control using Reduced-Order Observers”, in Proceedings of the 2008 American Control Confer-

ence, Seattle, WA, June 11–13, 2008.

18 of 18

American Institute of Aeronautics and Astronautics

You might also like