Вы находитесь на странице: 1из 17

Received: 20 November 2017 Revised: 21 January 2019 Accepted: 13 March 2019

DOI: 10.1002/rnc.4556

RESEARCH ARTICLE

High-order sliding-mode observer–based input-output


linearization

A. Ferreira de Loza1 L. Fridman3 L. T. Aguilar2 R. Iriarte3

1
Consejo Nacional de Ciencia y
Tecnología, Instituto Politécnico Summary
Nacional—Centro de Investigación y The problem of output control in multiple-input–multiple-output nonlinear sys-
Desarrollo de Tecnología Digital, Tijuana,
Mexico tems is addressed. A high-order sliding-mode observer is used to estimate the
2
Instituto Politécnico Nacional—Centro states of the system and identify the discrepancy between the nominal model
de Investigación y Desarrollo de and the real plant. The exact and finite-time estimation may be tackled as long
Tecnología Digital, Tijuana, Mexico
as the system presents the algebraic strong observability property. Thus, a con-
3
Facultad de Ingeniería, Universidad
Nacional Autónoma de México,
tinuous robust input-output linearization strategy can be obtained with respect
Mexico City, Mexico to a prescribed output. As a consequence, the closed-loop dynamics performs
robustly to uncertainties/perturbations. To illustrate the advantages of the pro-
Correspondence
L. T. Aguilar, Instituto Politécnico posed method, we introduce a study case that demands a robust linear system
Nacional—Centro de Investigación y behavior: the self-oscillations induced in an underactuated mechanical system
Desarrollo de Tecnología Digital,
through a two-relay controller. Experiments with an inertial wheel pendulum
22435 Tijuana, BCN, Mexico.
Email: laguilarb@ipn.mx illustrate the feasibility of the proposed approach.

Funding information K E Y WO R D S
Consejo Nacional de Ciencia y Tecnología,
disturbance identification, robust nonlinear control, sliding-mode control, uncertain systems
Grant/Award Number: 282013 and
285279; Programa de Apoyo a Proyectos
de Investigación e Innovación Tecnológica
of the Universidad Nacional Autónoma de
México, Grant/Award Number: IN115419

1 I N T RO DU CT ION

To control a system, system linearization is commonly considered around a given point, and then, linear control tech-
niques are applied to achieve the desired objective.1,2 Indeed, physical systems are intrinsically nonlinear, and, in some
cases, the effects of nonlinearity are determinant and cannot be safely disregarded. Over the past few years, nonlin-
earity effects have been tackled using diverse control strategies. Among them, input-output linearization methods offer
an approach to transform a nonlinear system into an equivalent linear system (see, eg, the works of Grizzle et al3 and
Shiriaev et al4 ). Nevertheless, two drawbacks of input-output linearization strategies should be emphasized: (i) their lack
of robustness against nonlinearities ignored in the system model and/or external disturbances that affect the system's
dynamics, and (ii) such techniques demand full-state measurements.
The present research focuses on using only output information to obtain a robust linear closed-loop behavior of the
system despite disturbances and neglected nonlinearities. Then, any linear control strategy can be applied.
A successful method for exact compensation against nonlinearities and disturbances is the sliding-mode control.5 In
practical situations, however, the robustness properties of the sliding-mode control are tarnished with the so-called chat-
tering phenomenon, ie, undesirable oscillations of finite frequency and amplitude. Higher-order sliding-mode (HOSM)
techniques maintain the robustness against uncertainties/disturbances while attenuating the chattering effect.6-8 The
aforementioned techniques, however, require full state feedback, which is not always available from measurements.

Int J Robust Nonlinear Control. 2019;1–17. wileyonlinelibrary.com/journal/rnc © 2019 John Wiley & Sons, Ltd. 1
2 FERREIRA DE LOZA ET AL.

Thus, an observer is needed. In this regard, several robust output controllers have been proposed in the literature
(see, for instance, other works9-15 and the references therein). The key issue of such techniques is the proposal of an
observer-controller scheme.
In the works of Peixoto et al9 and Oliveira et al,10 high-gain observers and sliding-mode controllers were brought
together to achieve the robust output feedback control for single-input–single-output disturbed nonlinear systems. Alter-
natively, in the works of Angulo et al12,13 and Oliveira et al,14 high-order sliding-mode observers (HOSMOs) were combined
with HOSM controllers. In the work of Angulo et al,12 the interconnection of HOSMOs and HOSM controllers pro-
duced a robust stabilization of multiple-input–multiple-output (MIMO) disturbed linear time-invariant (LTI) systems.
The method presented in the work of Angulo et al13 extends the results in another work of Angulo et al12 for MIMO nonlin-
ear disturbed systems. Moreover, the work of Angulo et al13 pursues the finite-time stabilization of the origin. Finite-time
stabilization of the origin is affordable only if the system under study is fully linearizable, which may be a restriction for
specific plants. The work of Oliveira et al14 goes a step further, proposing a global HOSM differentiator, which may be
combined with HOSM controllers to track the output of single-input–single-output uncertain systems with stable internal
dynamics.
The results published in the works of Peixoto et al,9 Oliveira et al,10 Fridman et al,11 Angulo et al,12,13 and Oliveira et al14
deal with nonlinear disturbed systems exploiting a discontinuous control action. Such discontinuous controllers aim
to counteract the effect of unknown but bounded uncertainties/disturbances. As a consequence, chattering may occur
when using the strategies proposed in the aforementioned works.9-14 The chattering effect is incompatible for certain
applications, for instance, in the orbital stabilization of mechanical systems,16 where the chattering effect spoils the qual-
ity of the obtained orbits. Indeed, in some applications, discontinuous controllers are capitalized for the generation of
self-oscillations.17 An alternative approach to ensure robustness against uncertainties/disturbances is, first, to construct
an observer that identifies such signals and, later, use the identified values to compensate for their effects by means of
a continuous control signal (see, for instance, the works of Fridman et al11 and Aguilar Ibañez et al15 ). In the work of
Aguilar Ibañez et al,15 a high-gain observer was proposed to estimate the unmeasured states as well as the uncertainties.
Later on, the control law used these estimations to compensate for the uncertainty effects through feedback. As a con-
sequence, the system was robustly stabilized without involving a discontinuous control action. The main drawback of
the work of Aguilar Ibañez et al15 is, however, that under the effect of time-varying bounded uncertainties, only practi-
cal stability may be attained. In the work of Fridman et al,11 the exact compensation of uncertainties is achieved using
observation and identification strategies via HOSM. The results in the work of Fridman et al11 are devoted to LTI systems
affected by uncertainties/disturbances.

1.1 Goal
The present research offers a continuous robust input-output exact linearization strategy for the control of MIMO
nonlinear uncertain systems.

1.2 Methodology
• A HOSMO estimates the states of the nonlinear system as well as the difference between the nominal model and the
real plant. Both state estimation and uncertainties identification are obtained, theoretically, in finite time. Thus, an
input-output exact linearization takes place.
• The robust convergence criterion presented in the work of Angulo et al12 is considered here to detect observer
convergence so that the controller may start at the right time until the transient time of the observer ends.

1.3 Contribution
The HOSMO-based input-output exact feedback linearization strategy provides a robust linear closed-loop system behav-
ior despite disregarded nonlinearities and/or disturbances. The main advantages of the proposed approach are the
following.
1. Only output information is considered.
2. Continuous robust input-output linearization is obtained using HOSM observation and identification.
3. A study case is presented in which the linearization strategy is combined with the two-relay control (TRC) to generate
robust exact self-oscillations around the open-loop unstable equilibrium point of an inertial wheel pendulum (IWP).
FERREIRA DE LOZA ET AL. 3

1.4 Organization
This paper is organized as follows. Section 2 presents the problem statement. Section 3 introduces the HOSMO that
estimates the full state and identifies the uncertainties together with external disturbances. Section 4 proposes the
input-output exact linearization controller. Section 5 summarizes the HOSMO-based input-output linearization method
in a step-by-step procedure. Section 6 combines the proposed strategy with the TRC method to robustify the generation
of self-oscillations in an IWP. Section 7 presents the experimental results that illustrate the effectiveness of the proposed
approach. Finally, Section 8 provides some concluding remarks.

1.5 Notation
The following notation is used. For any x ∈ ℜ, the symbol |x| denotes its absolute value. Let matrix A ∈ ℜn × m , with
rank(A) = m, the matrix A+ = (AT A)−1 AT is defined as the left pseudo-inverse of A. If x ∈ ℜn , then |x| stands for the
Euclidean norm. If A ∈ ℜn × m is a matrix, then the symbol |A| denotes the corresponding induced norm. The expression
1, p denotes the sequence of natural numbers 1, 2, … , p. Let h be a generic scalar function with a vector argument x
defined on an open set  ⊂ ℜn , h(x) ∶ ℜn → ℜ, define dh(x) = 𝜕h(x)∕𝜕x = [𝜕h(x)∕𝜕x1 𝜕h(x)∕𝜕x2 · · · 𝜕h(x)∕𝜕xn ]T . The
mapping 𝑓 (x) ∶  → ℜn is a vector field on , and the expression Lf h = 𝜕h∕𝜕x f (x) is called the Lie derivative of h along
the vector field f.

2 PROBLEM STAT EMENT

Consider a nonlinear time-invariant system


.
x = 𝑓 (x) + g(𝑦)(u(t) + 𝛾(t)) (1)

𝑦 = h(x) (2)

𝜎 = 𝜑(x), (3)
where x(t) = [x1 x2 … xn ]T is the state vector, u(t) ∈ ℜm is the control input, 𝛾(t) ∈ ℜm is the uncertainties/disturbances
vector representing the discrepancy between the nominal model and the real plant, y(t) ∈ ℜp is the measured output
vector with m ≤ p ≤ n, and t ≥ 0 is the time variable. The function 𝜎(t) ∈ ℜ corresponds to the output to be
controlled, and 𝜑(x) ∈ ℜ is a smooth function defined on an open set  ⊂ ℜn . The functions 𝑓 (x) ∶  → ℜn ,
g( y) = [ g1 ( y) g2 ( y) · · · gm ( y)], and h(x) = [h1 (x) h2 (x) · · · hp (x)]T are of appropriate dimensions. Let x∗ ∈  ⊂ ℜn be an
equilibrium point of the unperturbed system (ie, 𝛾(t) = 0 for all t ≥ 0).
Before continuing, some preliminary concepts are defined.
Definition 1 (Relative degree1 ).
Consider the system (1)-(2). The output y(t) ∈ ℜp is said to have a relative degree vector r = [r1 r2 … rn ]T with
respect to the input u(t) ∈ ℜm as well as to the disturbance vector 𝛾(t) ∈ ℜm , at the point x∗ ∈ ℜn if there exists a
neighborhood  of x∗ such that
(i)

Lg𝑗 L𝓁𝑓 h𝑗 (x) = 0, ∀𝑗 = 1, … , m, ∀𝓁 < ri − 1, ∀ i = 1, p


r −1
Lg𝑗 L𝑓i h𝑗 (x) ≠0 for at least one 1 ≤ 𝑗 ≤ m;

(ii) the p × m matrix


⎡ Lg1 L𝑓r1 −1 h1 (x) Lg2 L𝑓r1 −1 h1 (x) r −1
· · · Lgm L𝑓1 h1 (x) ⎤
⎢ r −1 r −1 r −1 ⎥
⎢ Lg L 2 h2 (x) Lg2 L𝑓2 h2 (x) · · · Lgm L𝑓2 h2 (x) ⎥
E(x) = ⎢ 1 𝑓 ⎥ (4)
⎢ ⋮ ⋮ ⋱ ⋮ ⎥
⎢ L Lrp −1 h (x) L Lrp −1 h (x) · · · Lgm L𝑓 hp (x) ⎥⎦
rp −1
⎣ g1 𝑓 p g2 𝑓 p

has full column rank.


4 FERREIRA DE LOZA ET AL.

Definition 2 (Algebraic strong observability13 ).


System (1)-(2) is said to be algebraically strongly observable with respect to the output y(t) ∈ ℜp , if a function F ∶
(k ) (k )
 → ℜn and integers ki , i = 1, p exist, such that x = F(𝑦1 , … , 𝑦1 1 , … , 𝑦p , … , 𝑦p p ).
The following assumptions are made in a neighborhood of the equilibrium point x∗ ∈  ⊂ ℜn .
Assumption 1. A constant 𝛾 + > 0 exists such that
.
|𝛾(t)| ≤ 𝛾 + , |𝛾(t)| ≤ 𝛾 + , ∀t ≥ 0. (5)

Assumption 2. The system (1)-(2) with a relative degree vector r = [r1 r2 … rn ]T at a point x∗ ∈  ⊂ ℜn has a total
∑p
relative degree i=1 ri = n. Therefore, system (1)-(2) is algebraically strongly observable.

Assumption 3. The full column rank matrix in (4) depends only on the output y, ie, E( y).

Assumption 4. For every initial condition x0 in a neighborhood of x∗ ∈  ⊂ n , for every control input u, and for
every 𝛾 satisfying Assumption 1, the solutions of system (1) are defined and remain upper bounded for all t ≥ 0.
Assumption 1 is realistic from a practical point of view. Assumption 2 expresses that the state and the disturbances can
be recovered in finite time. Concerning Assumption 3, the full-rank matrix in (4) may be rewritten in terms of the output
y as follows:
⎡ dL𝑓r1 −1 h1 (x) ⎤
⎢ r −1 ⎥
⎢ dL𝑓2 h2 (x) ⎥
E(𝑦) = ⎢ ⎥ g(𝑦); (6)
⎢ ⋮ ⎥
⎢ dLrp −1 h (x) ⎥
⎣ 𝑓 p ⎦
hence, it is a crucial assumption that allows withdrawing the control input effects on the observer error dynamics. Thus,
the controller may be switched on once the observer transient ends. Assumption 4 prevents finite-time escape and limits
the class of systems studied in this research to those that are forward complete.18 Assumptions 1 and 4 are instrumental
in the observer design since the local HOSM differentiator, first proposed in the work of Levant,19 is considered here.
Assumptions 1 and 4 can be relaxed using the global HOSM differentiator recently presented in the work of Oliveira et al.14
The control objective is to design a continuous HOSMO-based input-output exact linearization technique to guarantee
that the nonlinear plant (1) behaves as a linear one despite the discrepancies between the nominal model and the real
plant, as well as uncertainties/perturbations. The relevance of the method is illustrated in a study case. The case study
consists in exploiting the exact linearizing controller to robustify the self-oscillations generated by the two-relay controller
in a mechanical system. The goal is to obtain oscillations at the output 𝜎(t) with prescribed frequency and amplitude
regardless of whichever admissible disturbance 𝛾(t) affects the system.
To address this problem, in the following sections, we propose a HOSMO to estimate the state x(t) and identify the
perturbation 𝛾(t), theoretically in finite time. The observer design relies on a set of outputs that have a full relative degree
with respect to the uncertainties/disturbance signals, that is, satisfying the algebraic strong observability assumption, ie,
Assumption 2. Later, the estimated values are involved in the design of an input-output exact linearizing controller to
ensure that the original nonlinear system performs as a stabilizable linear system. In such circumstance, the output 𝜎(t)
may accomplish the prescribed task robustly, ie, self-oscillations via a two-relay controller in an IWP.
Opposite robust tracking closed-loop systems,20 the self-oscillating systems studied in the existing literature are sensitive
to external disturbances and parametric deviations, since the role of the two-relay controller is to induce oscillations at a
previously specified output 𝜎(t) ∈ ℜ, not to achieve a robust closed-loop system.17,21
The HOSMO, introduced to estimate the state vector and identify the uncertainties/disturbances in finite time, is
presented in the next section.

3 H O S M O D E SIG N

Here, the state vector will be recovered, and the uncertainties/disturbances vector will be identified. To this aim, first,
the system (1)-(2) will be transformed into the Brunovsky canonical form. Then, an auxiliary system will be proposed
FERREIRA DE LOZA ET AL. 5

to withdraw the control signal from the observation error dynamics. In consequence, the observation error dynamics is
represented as a perturbed chain of integrators. Finally, a HOSM differentiator will be applied at each coordinate of the
output error vector.
Notice that under Assumption 2, the state and perturbations can be recovered theoretically in finite time.
Let us represent (1)-(2) on a new basis
[ ]
r −1
𝜁 i = hi (x) L𝑓 hi (x) · · · L𝑓i hi (x) (7)

for i = 1, p. Thus, the following transformation


[ ]T
r −1
(x) = h1 (x) · · · L𝑓r1 −1 h1 (x) · · · hp (x) · · · L𝑓p hp (x) (8)

forms a local diffeomorfism. In other words, the inverse mapping


x = −1 (𝜁 ) (9)
exists in a neighborhood of the equilibrium point x∗ .
Now, taking into account (7) and the transformation (8), the system (1)-(2) can be rewritten in the following form:
.i
𝜁 𝑗 = 𝜁𝑗+1
i

.i r ( ) ∑
m
r −1 ( ) ( )
𝜁 ri = L𝑓i hi  (𝜁 ) +
−1
dL𝑓i hi −1 (𝜁 ) g𝑗 (𝑦) 𝛾𝑗 + u𝑗 (10)
𝑗=1

𝑦i = 𝜁1i
r −1
with i = 1, p, 𝑗 = 1, ri . Notice that the coefficients dL𝑓i hi (−1 (𝜁 ))g𝑗 (𝑦) are equal to the row elements in matrix E(y) in (6).
Now, the following auxiliary system is introduced to withdraw the control input from the observation error dynamics:
.i
𝜁̃ 𝑗 = 𝜁̃𝑗+1
.i ∑
m
r −1 ( )
𝜁̃ ri = dL𝑓i hi −1 (𝜁) g𝑗 (𝑦)u𝑗 (11)
𝑗=1

𝑦̃i = 𝜁̃1i

with i = 1, p, 𝑗 = 1, ri .
Let us introduce the observation error variable 𝜀 = 𝜁 − 𝜁̃ . From (10) and (11), the error dynamics follows as
.
𝜀i𝑗 = 𝜀𝑗+1
. r ( ) ∑m
r −1 ( )
𝜀iri = L𝑓i hi −1 (𝜁) + dL𝑓i hi −1 (𝜁 ) g𝑗 (𝑦)𝛾𝑗 (12)
𝑗=1

𝑦𝜀i = 𝜀i1

with i = 1, p, 𝑗 = 1, ri . Note that system (12) is a perturbed chain of integrators. Due to Assumption 2, system (10) and,
therefore, system (12) are algebraic strongly observable. Thus, given the outputs 𝑦𝜀i , the estimation of 𝜀i𝑗 (t) for all i = 1, p,
𝑗 = 1, ri can be achieved in finite time, differentiating the output 𝑦𝜀i , ri times.22 To this end, a HOSM differentiator is used
(see the work of Levant19 ).
The differentiator is given by
.
1
r +1 | |
ri
( )
zi0 = −𝜆i0 Γi i |z0i − 𝑦𝜀i | ri +1 sign z0i − 𝑦𝜀i + z1i
| |
.
1
r | . |
ri −1
( . )
zi1 = −𝜆i1 Γi i |z1i − zi0 | ri sign z1i − zi0 + z2i
| |
⋮ (13)
1 ( )
| |
1
. . .
ziri −1 = −𝜆iri −1 Γi2 |zri i−1 − ziri −2 | 2 sign zri i −1 − ziri −2 + zri i
| |
( )
.i i i .i
zri = −𝜆ri Γi sign zri − zri −1
6 FERREIRA DE LOZA ET AL.

for i = 1, p, where ri corresponds to the differentiator order and y𝜀i is the differentiator input. A positive sequence
(r )
{𝜆i1 , … , 𝜆iri } can be selected as in the work of Levant,19 and the gain Γi is a Lipschitz constant of 𝑦𝜀ii . Thus, under
Assumptions 1 and 4, there exists an a priori known set of constants Γ1 , … , Γp such that the inequality
| m ( |
| ri +1 ( −1 ) ∑ ( −1 ) ri −1 ( −1 ) . )|
|L hi  (𝜁 ) + L 2 ri
L h  (𝜁 ) 𝛾 (t) + L L h  (𝜁) 𝛾 (t) | < Γi (14)
| 𝑓 gi 𝑓 i i i gi 𝑓 i i |
| 𝑗=1 |
| |
holds for all t ≥ 0.
In the same spirit as the work of Rios et al,23 the latter makes sense from a physical point of view. For instance, in
mechanical systems, it is reasonable to assume an upper bound of the acceleration time derivative ( jerk), as well as an
upper bound for the initial estimation error.
Hence, a finite-time T > 0 exists such that the following holds:

𝜀̂ 11 = z01 , … , 𝜀̂ 1r1 = zr11 −1 , 𝜀̂ r1 = zr11


.1


, 𝜀̂ rp = zrp ,
p p p p .p p
𝜀̂ 1 = z0 , … , 𝜀̂ r1 = zr
p −1

where 𝜀̂ is the estimated value of 𝜀. Thus, 𝜀̂ ≡ 𝜀, for all t > T. Therefore, we have
𝜁̂ = 𝜀̂ + 𝜁,
̃ (15)
and the estimate of mapping (8) yields to the estimated value of 𝜁(t) denoted by 𝜁̂ = [𝜁̂ 1 𝜁̂ 2 … 𝜁̂ p ]T ∈ ℜn . Finally, the
exact estimate of x can be recovered from (9) as
̂
x̂ = −1 (𝜁). (16)
It is worth mentioning that the above inverse mapping is also local.

3.1 Uncertainties identification


Now, the finite-time estimates of 𝜀̂ ri for i = 1, p are available from the HOSM differentiator (13). Since (15), the estimated
.i

state 𝜁̂ is also accessible. Thus, consider the last row of the ith subsystems in (12), ie,

. r ( ) ∑m
r −1 ( )
𝜀iri = L𝑓i hi −1 (𝜁) + dL𝑓i hi −1 (𝜁) g𝑗 (𝑦)𝛾𝑗 (t)
𝑗=1

for all i = 1, p. Thus, the vector 𝛾(t) can be identified straightforwardly, ie,
( ( ))
𝛾̂ = E+ (𝑦) 𝜀̂ − Lr𝑓 h −1 (𝜁)
̂
.
, (17)

where E+ ( y) is the left pseudo-inverse matrix of E( y), 𝜀̂ = [𝜀̂ r1 𝜀̂ r2 … 𝜀̂ rp ]T , and Lr𝑓 h(−1 (𝜁))
̂ = [Lr1 h1 (−1 (𝜁))
̂ …
. .1 .2 .p
𝑓
rp −1 ̂ T
L𝑓 hp ( (𝜁 ))] , whereas 𝛾̂ ≡ 𝛾(t) for all t ≥ T.
Theorem 1. Consider system (1)-(2) and let Assumptions 1-4 hold. Then, the HOSMO (10)-(17) drives the estimated
states x̂ and identified signals 𝛾̂ toward the states x and the uncertainties/perturbations vector 𝛾, respectively, in finite time
T > 0, that is, limt→T ||x(t) − x̂ (t)|| = 0 and limt→T ||𝛾(t) − 𝛾̂ (t)|| = 0.

Proof. We split the proof into three parts.


First, system (1)-(2) is taken into a feedback linearizable form. Under Assumption 2, system (1)-(2) has a total rela-
∑p
tive degree, ie, i=1 ri = n. Hence, following the work of Isidori,1 the transformation (x) forms a local diffeomorfism.
As a consequence, the estimated trajectories of system (10) can be interpreted for the original system (1)-(2).
Second, the error dynamics is transformed into a perturbed chain of integrators in (12). Given that Assumption 3
holds, the coefficients multiplying the control signal uj , 𝑗 = 1, m, in the last row of (10), are entirely known. Thus, by
subtracting (11) from (10), the effects of the control input can be withdrawn from the error dynamics (12). Moreover,
under Assumptions 1 and 4, the error trajectories in (12) are bounded. As a consequence, the state of system (12) may
be recovered from the output 𝑦𝜀i and its successive ri derivatives employing a HOSM differentiator (13), and a suitable
gain Γi following (14), for every i = 1, p.
FERREIRA DE LOZA ET AL. 7

For the last part of the proof, the finite-time convergence of the estimated state and the identified uncertainty/
perturbation signal are stated. With the proper choice of the constants 𝜆i𝑗 as shown in the work of Levant19 and
choosing Γi in agreement with (14), then a finite-time T > 0 exists such that

z𝑗i = d𝑗 𝑦𝜀i ∕dti

is fulfilled for all 𝑗 = 0, ri , i = 1, p. Accordingly, since 𝜀 = 𝜁 − 𝜁̃ , the estimated vector 𝜁̂ is recovered by algebraic
manipulation in (15). Therefore, the estimated state x̂ is plainly recovered in (16).
Concerning the identification of the uncertainties/disturbances vector 𝛾, which represents the difference between
the nominal model and the real system, provided that under Assumption 1, matrix E( y) has full rank, 𝛾̂ can be obtained
forthright in (17). The proof of Theorem 1 is complete.

It is worth noticing that the HOSM differentiator in (13) gives the best accuracy order with respect to bounded deter-
ministic noise and discretization issues.19 The accuracy analysis for the proposed HOSMO comes straightforwardly from
the results in the work of Angulo et al.12
Proposition 1. Let system (2) satisfy Assumptions 1 and 4. Thus, for any sufficiently large Γi > 0 following (14), the
HOSM ri th-order differentiators in (13) with proper parameters {𝜆i1 , … , 𝜆iri } with i = 1, p assure that z𝑗i = d𝑗 𝑦𝜀i ∕dt𝑗
after a finite-time t > T. Assume that the outputs 𝑦𝜀i are affected with measurement noise uniformly bounded by |n(t)| <
𝜅n Γi 𝛿 ri +1 , where 𝛿 is a positive parameter, and 𝜅 n > 0 is a constant. Besides, assume 𝑦𝜀i are sampled with a sampling
step 𝜏 > 𝜅 𝜏 𝛿, 𝜅 𝜏 > 0 is a constant. Then, the following accuracies
| i | ( )
|z𝑗 − d𝑗 𝑦𝜀i ∕dt𝑗 | = O 𝛿 ri +1−𝑗 , 𝑗 = 0, ri , i = 1, p (18)
| |
hold after a finite-time transient.

Remark 1 (Separation principle).


Before introducing the controller, it is worth noticing that reliable state estimation and disturbance identification
are available after a finite-time transient. Therefore, the controller turns on once the transient ends. Several works
devoted to the output-feedback control of nonlinear systems10-13 consider such time-dwell strategy. The main reason
for applying a time-dwell strategy is to prevent the finite-time escape of the closed-loop system trajectories. In the
present work, we exploit a criterion specifically proposed for HOSMOs, first presented in the work of Angulo et al.12
Such criterion consists in verifying the differences |z0i − 𝑦𝜀i | = O(𝛿 ri +1 ) for i = 1, p to detect the moment when the
HOSM differentiator converges. Therefore, the convergence time T may be estimated once and forever by simulation.

4 INPUT- OUTPUT EXACT LINEARIZATION CO NTRO LLER

Let assume that the observer transient ended; therefore, the equalities x ≡ x̂ and 𝛾 ≡ 𝛾̂ are certainly obtained. Hence,
for the sake of clarity, the control output, which possibly depends on the estimated state x̂ , will be simply represented as
𝜎 = 𝜑(x). Hereinafter, the following is assumed.
Assumption 5. The output 𝜎 ∈ ℜ has a total relative degree r̄ ≤ n at a point x∗ ∈  ⊂ ℜn , and the internal
dynamics of (1) and (3) is asymptotically stable.

Assumption 6. Let the vector E(x) ̄ = [ Lg1 Lr𝑓̄ −1 𝜑(x) · · · Lgm Lr𝑓̄ −1 ]. Thus, there exist at least one component such that
Lgi Lr𝑓̄ −1 𝜑(x) ≠ 0, with 1 ≤ i ≤ m in a neighborhood of x∗ ∈ Ω ⊂ ℜn .

Assumption 7. System (1) has an involutive distribution

G = span {g1 (𝑦), g2 (𝑦), … , gm (𝑦)}

in a neighborhood of x∗ ∈  ⊂ ℜn .
The aforementioned Assumptions establish the conditions for attaining the input-output linearization of system (1),
with respect to output 𝜎 in (3).
8 FERREIRA DE LOZA ET AL.

Thus, systems (1) and (3) may be represented considering the new basis
[ ]
𝜂 = 𝜑(x) L𝑓 𝜑(x) · · · Lr𝑓̄ −1 𝜑(x) (19)

with 𝜂 ∈ ℜr̄ . Then, there exist n − r̄ functions 𝜂,


̆ such that the map
[ ]T
 (x) = 𝜂(x)
̆ · · · 𝜂̆ n−̄r (x) 𝜑(x) · · · Lr𝑓̄ −1 𝜑(x) (20)

with  ∈ ℝn forms a local diffeomorfism in a neighborhood of x∗ ∈  ⊂ ℜn with Lg𝑗 𝜂̆ i = 0 for all i = 1, n − r̄ , 𝑗 = 1, m.


Therefore, the map (20) transforms systems (1) and (3) into
·
𝜂̆ = 𝑓0 (𝜂,
̆ 𝜂)
.
𝜂 𝑗 = 𝜂𝑗+1
. ( ) ∑
m
r̄ −1 ( ) (21)
𝜂 r̄ = Lr𝑓̄ 𝜑  −1 (𝜂, 𝜂)
̆ + Lg𝑗 L𝑓i 𝜑i  −1 (𝜂, 𝜂)
̆ (ui (t) + 𝛾i (t))
𝑗=1

𝜎 = 𝜂1

with 𝑗 = 1, r̄ − 1, 𝜂 = [𝜑(x) … Lr𝑓̄ −1 𝜑(x)]T ∈ ℜr̄ and 𝜂̆ ∈ ℜn−̄r , 𝜂̆ = [𝜂̆ 1 … 𝜂̆ n−̄r ]T . Notice that the coefficients
r̄ −1 .
Lg𝑗 L𝑓i 𝜑i ( −1 (𝜂, 𝜂) ̄
̆ are exactly equal to the column elements of E(x). ̆ 𝜂) = 𝜕𝜕x𝜂̆ 𝑓 ( −1 (x)) represents the
The term 𝜂̆ = 𝑓0 (𝜂,
internal dynamics.
The following control may be proposed:
( ) ( r̄ ( −1 ) )
u = −Ē †  −1 (𝜂, 𝜂)
̆ L𝑓 𝜑  (𝜂, 𝜂)
̆ + K𝜂 + 𝜈 − 𝛾̂ (t), (22)

where u = [u1 … um ]T , Ē † (x) is the right-inverse defined by E† = Ē T (x)(E(x)


̄ Ē T (x))−1 , and it exists in a neighborhood of
x ∈  ⊂ ℜ . The
∗ n
[ variable ]𝛾̂ (t) ∈ ℜ is the identified value of the uncertainties/disturbances vector. The matrix gain
m

K ∈ ℜ1×̄r , K = k1 … kr̄ , with kj > 0 for every 𝑗 = 1, r̄ is a stabilizing gain. Finally, 𝜈 ∈ ℜ is the control input
appointed to attain the desired closed-loop behavior of the linearized system.
The following theorem states the results mentioned above.
Theorem 2. Let Assumptions 4-7 be satisfied. Thus, systems (1) and (3) with the control law (22) become a linear
stabilizable system in spite of the perturbations 𝛾(t).

Proof. We will analyze the stability of the closed-loop system (21)-(22).


Under Assumption 5, systems (1) and (3) have a relative degree r̄ ≤ n. Since r̄ < n, due to Assumption 7, it is always
possible to find 𝜂̆ i functions for i = 1, … , r̄ satisfying Lg𝑗 𝜑(x) ≠ 0 for every x ∈  ⊂ ℜn . Thus, following the work
of Isidori,1 the transformation  (x) given in (20) forms a local diffeomorfism in a neighborhood of x∗ ∈  ⊂ ℜn .
Therefore, system (1)-(3) can be expressed as in (21). Second, substituting (22) into (21) yields
.
𝜂̆ = 𝑓0 (𝜂, 𝜂)
̆
.
𝜂 = A𝜂 + B𝜈 (23)
𝜎 = C𝜂.
The triplet (A, B, C) is given by

⎡ 0 1 0 ··· 0 ⎤ ⎡0⎤
⎢ 0 0 1 ··· 0 ⎥ ⎢0⎥ [ ]
A=⎢ , B = ⎢ ⎥, C = 1 0 ··· 0 .
⎢ ⋮ ⋮ ⋮ ⋱ ⋮ ⎥⎥ ⎢⋮⎥
⎣ −k1 −k2 −k3 ··· −kr̄ ⎦ ⎣1⎦

Thus, from the above equations, it follows that the pair (A, B) is controllable. Moreover, given that ki > 0, i = 1, r̄ ,
then A is Hurwitz. Besides, for 𝜈 = 0, the vector 𝜂 tends to the origin; thus, under Assumption 5, the variable 𝜂̆ will
be asymptotically stable. This completes the proof of Theorem 2.

In the next section, we recapitulate the robust input-output exact linearization method in a step-by-step procedure.
FERREIRA DE LOZA ET AL. 9

5 S U M M A RY O F T H E METH O D

Consider the nonlinear system (1)-(3) satisfying Assumptions 1-7. The robust input-output exact linearization proce-
dure can be synthesized in four steps: (1) transformation of the observation error into a perturbed chain of integrators,
(2) HOSM differentiation process, (3) recovery of the estimated state and the identified value of the uncertainties vector
signal, and (4) input-output exact linearization control law computation. The steps are illustrated in Figure 1.
Step 1) Transformation of the observation error into a perturbed chain of integrators. First, transform (1)-(2) into the new
coordinates (7). Then, construct the auxiliary system (11). Two variables are recovered in this step: the auxiliary
state 𝜁̃ = [ 𝜁̃11 · · · 𝜁̃r11 · · · 𝜁̃1 · · ·𝜁̃rp ]T and the observation error 𝑦𝜀i = 𝑦i − 𝑦̃i for i = 1, p.
p p

Step 2) Differentiation processes. Following (13), design p-HOSM differentiators of ri th order each one, with i = 1, p.
Differentiate by ri times every 𝑦𝜀i , i = 1, p. The vectors 𝜀̂ = [ z01 · · · zr11 −1 · · · z0 · · · zrp −1 ]T and 𝜀̂ = [ zr11 · · · zrp ]T
p p . p

are retrieved in this step.


Step 3) Recovery of the estimated state vector x̂ as well as the identified uncertainty vector 𝛾̂ . The exact state estimate x̂
is retrieved from (15)-(16) combining 𝜀̂ and 𝜁, ̃ whereas the identified value of the uncertainties/perturbations
vector 𝛾̂ (t) is obtained straightforwardly from (17).
Step 4) Robust input-output exact linearization controller. Finally, the control law (22) is computed based on the estimated
state x̂ and the identified signal 𝛾̂ .
In the sequence, we apply the HOSMO-based input-output linearization technique in a case of study: robust generation
of self-oscillations in an IWP using the TRC.
Before continuing, it is worth mentioning that relay-type nonlinearity is a discontinuous nonlinearity. The effect of
discontinuous control actions, for instance, may provoke such nonlinearity. The relay-type nonlinearities tarnish the
controlled output with chattering. The chattering phenomenon is an undesirable oscillation whose effects and attenuation
strategies have been investigated (see, for instance, the work of Kamachkin et al24 ).
The TRC, on the contrary, has been exploited to generate oscillations of prescribed frequency and amplitude in mechan-
ical systems.17 The main constraint, however, is that the TRC method is highly sensitive to uncertainties/disturbances.

FIGURE 1 Summary of the high-order sliding-mode observer–based input-output linearization method: Step 1) Observation error
transformation; Step 2) Higher-order sliding-mode differentiation processes; Step 3) Recovery of the estimated state x̂ and the identified
uncertainty vector 𝛾̂ ; Step 4) Computation of the robust input-output exact linearization control law u [Colour figure can be viewed at
wileyonlinelibrary.com]
10 FERREIRA DE LOZA ET AL.

In other words, TRC requires the system to perform with a nominal linear behavior. Thus, in the next section, the robust
input-output exact linearization method presented here will be instrumental in generating robust oscillations at the
underactuated coordinate of an IWP.
Remark 2. Although the techniques introduced in other works9-14 guarantee robustness against disturbances, they
cannot be straightforwardly combined with TRC. The discontinuous nature of the controllers previously proposed9-14
may provoke chattering, which is incompatible with the TRC goal.

6 C A S E O F ST U DY: RO B U ST G ENERATION OF SELF- OSCILLATIONS IN AN


IWP USING TRC

The dynamics of an IWP, taken from the work of Block et al,25 augmented with uncertainties, corresponding to disregarded
system dynamics, are given by
⎡ x3 ⎤ ⎡ 0 ⎤
⎢ x4 ⎥ ⎢ 0 ⎥
. ⎢ ⎥ ⎢ ⎥
⎥ + ⎢ − J −J ⎥(u + 𝛾),
x = ⎢− 1 1 (24)
h sin x 1
⎢ J1 −J2 ⎥ ⎢ 2 −1 1 ⎥
⎢ 1 h sin x1 ⎥ ⎢ J1 J2 ⎥
⎣ J1 −J2 ⎦ ⎣ J2 −J1 ⎦
⏟⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏟ ⏟⏞⏞⏟⏞⏞⏟
𝑓 (x) g(x)
𝑦 = [x1 x2 ]T , (25)
( )
𝜎 = J1 x3 + J2 x4 + k x1 − 𝜋 + J1−1 J2 x2 (26)
with x = [x1 x2 x3 x4 ] , where x1 (t) ∈ ℜ is the angle of the pendulum, counted clockwise from the vertical downward
T

position; x2 (t) ∈ ℜ is the absolute angle of the disk; x3 (t) ∈ ℜ corresponds to the pendulum angular velocity; x4 (t) ∈ ℜ
stands for disk angular velocity; J1 , J2 , h, and k are positive physical parameters that depend on the geometric dimensions
and the inertia-mass distribution; u(t) ∈ ℜ is the adjustable torque applied to the disk; and 𝛾(t) ∈ ℜ represents the
uncertainties affecting the system such as friction forces acting in the actuator. The measured output is represented by
y(t) ∈ ℜ2 . A schematic representation of the IWP is given in Figure 2.
Two goals are pursued: first, the robust stabilization of the pendulum in its unstable equilibrium point and, second, the
generation of self-oscillations at the scalar output 𝜎(x) defined in (26) around the upright position (x1∗ , x3∗ ) = (𝜋, 0) of the
uncertain IWP (24). In both cases, only position measurements are considered for feedback.

6.1 HOSMO-based input-output linearization of the IWP


Given that IWP is a mechanical system, Assumption 1 is granted. Assumption 2 is also satisfied, with r1 = r2 = 2 and a
total relative degree r = 4. By direct computation in (6), matrix E( y) yields to
[ J
]T
1
E(𝑦) = − J −J J (J 1−J ) , (27)
1 2 2 1 2

which has full row rank, fulfilling Assumption 3.

FIGURE 2 Inertial wheel pendulum


FERREIRA DE LOZA ET AL. 11

Therefore, system (24)-(26) is algebraically strongly observable. In consequence, the state vector and the perturbations
can be estimated in finite time by employing a HOSMO.
Thus, system (24) can be represented on a new basis by means of the coordinates transformation 𝜁11 = x1 , 𝜁21 = x3 ,
𝜁1 = x2 , 𝜁22 = x4 , ie,
2
.1
𝜁 1 = 𝜁21
.1
𝜁 2 = (J2 − J1 )−1 h sin 𝜁11 + (J2 − J1 )−1 (𝛾 + u)
.2 (28)
𝜁 1 = 𝜁22
.2
𝜁 2 = J1 J2−1 (J2 − J1 )−1 h sin 𝜁11 + (J2 − J1 )−1 (𝛾 + u),
with 𝑦 = [𝜁11 𝜁12 ]T .
Step 1) By virtue of (11), the following auxiliary systems are proposed to transform the observation error into a perturbed
chain of integrators:
.1
𝜁̃ 1 = 𝜁̃ 1 2
.1
𝜁̃ 2 = (J2 − J1 )−1 u
.2
𝜁̃ 1 = 𝜁̃22
.2
𝜁̃ 2 = J1 J2−1 (J2 − J1 )−1 u,
with 𝑦̃ = [ 𝜁̃11 𝜁̃12 ]T . Thus, we retrieve the next information
[ ]T
𝜁̃ = 𝜁̃11 𝜁̃21 𝜁̃12 𝜁̃22 , 𝑦𝜀 = 𝑦 − 𝑦.
̃ (29)

Step 2) From (13), design two HOSM differentiators


1 2 ( )
z10 = −𝜆10 Γ13 ||z01 − 𝜀11 || 3 sign z01 − 𝜖11 + z11
.

.
1
| . |
1
( . )
z11 = −𝜆11 Γ12 |z11 − z10 | 2 sign z11 − z10 + z21
| |
. ( . )
z12 = −𝜆12 Γsign z21 − z11
1
| |
2
( )
z02 = −𝜆20 Γ23 |z02 − 𝜀21 | 3 sign z02 − 𝜀21 + z12
| |
1
| . |
1
( . )
z12 = −𝜆21 Γ22 |z12 − z20 | 2 sign z12 − z20 + z22
| |
( . )
z22 = −𝜆22 Γ2 sign z22 − z21 ,
redeeming, as a result, the following information:
[ ]T . [ ]T
𝜀̂ = z01 z11 z02 z22 , 𝜀̂ = z21 z22 . (30)

Step 3) Use 𝜁̃ and 𝜀̂ retrieved in (29) and combine them in (15) to obtain 𝜁̂ . Later on, apply the inverse map (16) to recover
the estimated state
[ ]T
x̂ = 𝜁̂11 𝜁̂12 𝜁̂21 𝜁̂22 . (31)

Now, we obtain 𝛾̂ (t) by substituting 𝜀, E( y), and L4𝑓 h(̂x) = [ (J2 − J1 ) h sin(x1 ) J1 J2 (J2 − J1 ) h sin(x1 ) ]T in (17).
−1 −1 −1

In view of Theorem 1, the expressions x̂ ≡ x and 𝛾̂ ≡ 𝛾 are certainly obtained for all t ≥ T. It is worth mentioning
that such time T may be computed by simulation once and forever by virtue of the results presented in the work
of Angulo et al.12
Step 4) Considering (26) as the output to be controlled, an input-output linearization with local stability of the zero
dynamics can be achieved for the IWP (see the work of Grizzle et al3 for further details). The output 𝜎 has a
relative degree r̄ = 3. A new basis (19) is proposed as
𝜂1 = 𝜎,
. .
𝜂2 = kJ1−1 J2 x2 + kx1 − h sin(x1 ),
.
𝜂3 = −kJ1−1 h sin(x1 ) − h cos(x1 )x1 .
12 FERREIRA DE LOZA ET AL.

Moreover, the function


𝜂̆ 1 = x1 − 𝜋 + J1−1 J2 x2

satisfies 𝜕 𝜂̆ 1 g(x)∕𝜕x = 0. Thus, the map [𝜂1 𝜂2 𝜂3 𝜂̆ 1 ]T forms a local diffeomorfism in a neighborhood of the equilibrium
.
point [x1∗ x1 ] = [𝜋 0]. System (24) can be rewritten as
.
𝜂 1 = 𝜂2
.
𝜂 1 = 𝜂3
( ) ( )
̆ + Ē  −1 (𝜂, 𝜂)
.
𝜂 3 = L3𝑓 𝜑  −1 (𝜂, 𝜂) ̆ (u + 𝛾) (32)
.
𝜂̆ 1 = −k𝜂̆ 1 + J1−1 𝜂1
𝜎 = 𝜂1 ,
where
( ) h cos(x1 )
Ē  −1 (𝜂, 𝜂)
̆ =
J1 − J2
( ) kh h2
L3𝑓 𝜑  −1 (𝜂, 𝜂)
̆ = hx32 sin(x1 ) − x3 cos(x1 ) + cos(x1 ) sin x1 .
J1 J1 − J2
.
The zero dynamics, 𝜂̆ = 𝑓0 (𝜂, 𝜂)
̆ = −k𝜂̆ 1 + J1−1 𝜂1 , is asymptotically stable. Hence, Assumption 5 is satisfied.
Designing u in accordance with (22) yields
( )
J1 − J2 k h
u=− 2
x sin(x1 ) − x3 cos(x1 ) + cos(x1 ) sin x1 − K𝜂 − 𝛾̂ (t), (33)
cos(x1 ) 3 J1 J1 − J2
[ ]
where K = k1 k2 k3 , with k1 , k2 , k3 > 0.
Finally, substituting (33) into (32) yields the following linear closed-loop dynamics:
⎡ 0 1 0 ⎤ ⎡0⎤
𝜂 = ⎢ 0 0 1 ⎥𝜂 + ⎢ 0 ⎥ 𝜈,
.
(34)
⎢ ⎥ ⎢ ⎥
⎣ −k1 −k2 −k3 ⎦ ⎣1⎦
⏟⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏟ ⏟⏟⏟
 B
. k 1 [ ]
𝜂̆ 1 = − 𝜂̆ 1 + 𝜎, 𝜎 = 1 0 0 𝜂.
J1 J1 ⏟⏞⏟⏞⏟
C

In accordance with Theorem 2, the closed-loop dynamics of the IWP in (34) possess a nominal linear behavior despite
the uncertainties/perturbations 𝛾(t). Therefore, the control input 𝜈 will be designed using the TRC technique. The TRC
methodology is summarized in the following subsection (see the work of Aguilar et al17 for further details).

6.2 TRC to induce self-oscillations in the IWP


Let us consider the linearized plant (23), that is,
. ̂ + B𝜈
𝜂 = A𝜂
(35)
𝜎 = C𝜂
̂ B, C defined in (34), and 𝜈 ∈ ℜ is the TRC input given by
with matrices A,
.
𝜈 = −c1 sign(𝜎) − c2 sign(𝜎), (36)
where c1 and c2 are design parameters. The proper selection of c1 and c2 may provoke a steady periodic motion with a
desired frequency and amplitude at the scalar output 𝜎.
Three methods have been published to compute TRC gains: the describing function (DF), the locus of a perturbed relay
system (LPRS), and the Poincaré map. The Poincaré map is the most accurate method to tune TRC gains; however, its
analytical solution cannot be straightforwardly obtained. The LPRS is based on the approximate DF method but offers
an exact analysis of the oscillatory properties of relay feedback systems.26 Therefore, the method used in this paper will
be LPRS.
FERREIRA DE LOZA ET AL. 13

x 10
0.5

0.0014
0

−0.5

Imag J( )
−1

−1.5

J( )
−2

−0.0022

−2.5
−1 −0.5 0 0.5 1 1.5
Real J( ) x 10

FIGURE 3 The locus of a perturbed relay system for the linearized model of the inertial wheel pendulum

The LPRS is the characteristic of the response of a linear plant to an unequally spaced pulse control of variable frequency
in a closed-loop system. LPRS requires a significant computational effort, but it yields an exact solution (see the work of
Aguilar et al17 and the references therein for further details).
Given a desired frequency Ω and amplitude , the following general formulas, derived from the LPRS method in the
work of Aguilar et al,17 are introduced:
𝜋  1
c1 ≈ √ , (37)
4 |W( 𝑗Ω)| 1 + 2𝜉 cos(2𝜋𝜃) + 𝜉 2
𝜋  𝜉
c2 ≈ √ , (38)
4 |W( 𝑗Ω)| 1 + 2𝜉 cos(2𝜋𝜃) + 𝜉 2
to obtain c1 and c2 . Here, 𝜃 ∈ [ −1∕2, 1∕2] is the value of the time shift between the switching of the first and second
.
relays that can be found from 𝜎(0) = 0, 𝜉 is the ratio between the DF of the second relay and the first relay, and W( jΩ) is
the corresponding frequency domain transformation of (35), that is, W(s) = C(sI − A) ̂ −1 B, with s = jΩ.
As regards the computation of TRC gains c1 and c2 for IWP, first, the corresponding transfer function of the output
linearized model in (34) is
1
W(s) = 3 . (39)
s + k3 s2 + k2 s + k1
Figure 3 illustrates the locus of a perturbed two-relay system (LPRS) for the linearized model of the IWP (34) and highlights
the intersection of  (Ω) = 0.0014 − 𝑗0.0022 and  (𝜔) locus, where 𝜔 ∈ [0, ∞). According to the work of Boiko,26 the
frequency domain function  (𝜔) is defined as a characteristic of the response of the linear part to the unequally spaced
pulse input 𝜈(t), subject to the reference input as 𝜔 varies. The amplitude of the attained oscillation in x1 , denoted as x1 ,
is given by √
 ( )−1
x1 ≈ 1 + J12 Ω2 + k2 . (40)
h

In the work of Aguilar et al,21 it was shown that the closed-loop system (34), controlled by the two-relay algorithm (36)
with the parameters c1 and c2 selected as in (37) and (38), exhibits a periodic trajectory 𝜂 ∗ . The periodic trajectory 𝜂 ∗ has
a desired amplitude  and frequency Ω; moreover, it is orbitally exponentially stable.

7 EXPERIMENTAL RESULTS

In this section, we present the experimental results for the IWP manufactured by Quanser Inc, shown in Figure 2. The
experimental setup includes a PC equipped with a real-time dSPACE acquisition platform with a minimum sample and
14 FERREIRA DE LOZA ET AL.

TABLE 1 Closed-loop experimental parameters


J1 4.572 × 10−3 K [ 22 155 350 ]
J2 2.495 × 10−5 k 1 × 10−4
h 0.4594 x1 0.05
Γ1 8.5 × 103 Ω 12 rad/s
Γ2 17 × 106 c1 2
𝜆0, … ,2 {1.1, 1.5, 2} c2 −2.5

FIGURE 4 Synthesis of the robust self-oscillation generation procedure [Colour figure can be viewed at wileyonlinelibrary.com]

integration time equal to 50 μs. The pendulum and wheel angles are measured by two encoders with a resolution of 1024
and 500 pulses per rate, respectively. The parameters used for the IWP closed-loop experiment are listed in Table 1.
Figure 4 shows a block diagram that summarizes the overall system. Starting from the nonlinear model of the IWP
affected by uncertainties/disturbances in (24) and from the measured set of outputs (25), a HOSMO is designed to esti-
mate the state x̂ and uncertainties 𝛾̂ . Those values are involved in the synthesis of a linearization control law in (33).
In consequence, the nonlinear plant acts as a nominal linear system with stable internal dynamics (34). Finally, the
TRC controller (36)-(38) is injected to the plant to induce the self-oscillation of desired frequency and amplitude at the
output 𝜎.
Two scenarios are studied: first, the robust stabilization of the underactuated coordinate x1 at its unstable equilibrium
point, stage (a); second, the TRC is applied to generate oscillations of prescribed frequency and amplitude, stage (b). To
evidence the robustness of the proposed methodology, both scenarios are contrasted with a conventional input-output
exact linearization approach (see, for instance, the work of Isidori1 ). It should be noted that the input-output linearization
technique in the work of Isidori1 relies on perfect knowledge of the model.
(a) Robust stabilization of x1 (t). Figure 5 depicts the stabilization of the underactuated coordinate x1 . In the time interval
15 [s] ≤ t < 32 [s], a conventional linearization controller1 is applied. Due to uncertainties/disturbances effects, x1
oscillates around the equilibrium point. By contrast, for t ≥ 32 [s], the HOSMO-based input-output linearization
controller given by (33) is used; in consequence, x1 reaches its unstable equilibrium point and remains there.
(b) Robust oscillations of x1 (t). Figure 6 shows the oscillation of the underactuated coordinate x1 as well as the time
evolution of the estimated coordinate x2 . For the sake of comparison, Figure 7 contrasts the performance of the
HOSMO-based input-output linearizing controller against the results presented in the work of Aguilar et al.21 Clearly,
the oscillation quality is improved when our approach is used, as shown on the shadowed side of the Figure.
Let us recall that x2 is a cyclic variable corresponding to the absolute angle of the disk; therefore, its growth does not
imply that the closed-loop system is unstable (see the work of Grizzle et al.3 ).
FERREIRA DE LOZA ET AL. 15

FIGURE 5 Stabilization of underactuated variable x1 (t). The uncertainties provoke an undesired oscillation of x1 (t) around its open-loop
unstable equilibrium point. By contrast, x1 reaches its open-loop unstable equilibrium point and remains there ∀t ≥ 32 [s]. The time
evolution of coordinate x̂ 2 is also shown [Colour figure can be viewed at wileyonlinelibrary.com]

FIGURE 6 Robust oscillation of x1 and the time evolution of coordinate x2 [Colour figure can be viewed at wileyonlinelibrary.com]

FIGURE 7 Comparison of x1 oscillations using the high-order sliding-mode observer–based input-output linearization (shadowed side)
versus the approach used in the work of Aguilar et al.21 The time evolution of coordinate x2 is also depicted [Colour figure can be viewed at
wileyonlinelibrary.com]
16 FERREIRA DE LOZA ET AL.

8 CO N C LUSION S

The problem of robust input-output exact linearization control for nonlinear uncertain systems has been addressed. A
HOSMO was used to estimate the states of the system and identify the discrepancies between the nominal model and
the real plant, theoretically in finite time. Thus, a robust input-output linearization controller was proposed for a class
of minimum-phase systems. The linearization method presented herein is instrumental in applications sensitive to the
chattering effect, such as the TRC. To illustrate this issue, the linearization approach was combined with TRC for the
generation of robust self-oscillations in an IWP. Experimental results demonstrated the feasibility of the strategy developed
in this paper.

ACKNOWLEDGEMENTS
The authors gratefully acknowledge the financial support from the Consejo Nacional de Ciencia y Tecnología under grants
282013 and 285279 and the Programa de Apoyo a Proyectos de Investigación e Innovación Tecnológica of the Universidad
Nacional Autónoma de México under grant IN115419. Maria Isabel Perez Montfort corrected the English version of this
paper.

ORCID

A. Ferreira de Loza https://orcid.org/0000-0003-1189-0857


L. Fridman https://orcid.org/0000-0003-0208-3615
L. T. Aguilar https://orcid.org/0000-0001-6603-6485

REFERENCES
1. Isidori A. Nonlinear Control Systems. London, England: Springer-Verlag; 1996.
2. Yang K, Ji H. Output feedback hierarchical control for nonlinear systems. Int J Robust Nonlinear Control. 2017;27(17):4089-4103.
3. Grizzle J, Moog C, Chevallereau C. Nonlinear control of mechanical systems with an underactuated cyclic variable. IEEE Trans Autom
Control. 2005;50:559-576.
4. Shiriaev AS, Freidovich LB, Robertsson A, Sandberg A. Virtual-holonomic-constraints-based design of stable oscillations of Furuta
pendulum: theory and experiments. IEEE Trans Robotics. 2007;23:827-832.
5. Utkin VI. Sliding Modes in Control and Optimization. Berlin, Germany: Springer-Verlag; 1992.
6. Bartolini G, Ferrara A, Usai E. Chattering avoidance by second-order sliding mode control. IEEE Trans Autom Control. 1998;43(2):241-246.
7. Raimondo DM, Rubagotti M, Jones CN, Magni L, Ferrara A, Morari M. Multirate sliding mode disturbance compensation for model
predictive control. Int J Robust Nonlinear Control. 2015;25(16):2984-3003.
8. Pérez Ventura U, Fridman L. When is it reasonable to implement the discontinuous sliding-mode controllers instead of the continuous
ones? Frequency domain criteria. Int J Robust Nonlinear Control. 2019;21(3):810-828.
9. Peixoto AJ, Oliveira TR, Hsu L, Lizarralde F, Costa RR. Global tracking sliding mode control for a class of nonlinear systems via variable
gain observer. Int J Robust Nonlinear Control. 2011;21(2):177-196.
10. Oliveira TR, Peixoto AJ, Hsu L. Peaking free output-feedback exact tracking of uncertain nonlinear systems via dwell-time and norm
observers. Int J Robust Nonlinear Control. 2013;23(5):483-513.
11. Fridman L, Ferreira de Loza A, Aguilar LT. Robust output control of systems subjected to perturbations via high-order sliding modes
observation and identification. In: Fridman L, Barbot JP, Plestan F, eds. Recent Trends in Sliding Mode Control. 2016:57-76.
12. Angulo MT, Fridman L, Levant A. Output-feedback finite-time stabilization of disturbed LTI systems. Automatica. 2012;48:606-611.
13. Angulo MT, Fridman L, Moreno JA. Output-feedback finite-time stabilization of disturbed feedback linearizable nonlinear systems.
Automatica. 2013;49(9):2767-2773.
14. Oliveira TR, Estrada A, Fridman L. Global and exact HOSM differentiator with dynamic gains for output-feedback sliding mode control.
Automatica. 2017;81(7):156-163.
15. Aguilar Ibañez C, Sira Ramirez H, Acosta JA. Stability of active disturbance rejection control for uncertain systems: a Lyapunov
perspective. Int J Robust Nonlinear Control. 2017;27(18):4541-4553.
16. Pchelkin S, Shiriaev A, Robertsson A, Kolyubin S, Paramonov L, Gusev S. On orbital stabilization for industrial manipulators: case study
in evaluating performances of modified PD and inverse dynamics controllers. IEEE Trans Control Syst Tech. 2017;25(1):101-117.
17. Aguilar L, Boiko I, Fridman L, Iriarte R. Self-Oscillations in Dynamic Systems: A New Methodology via Two-Relay Controllers. London, UK:
Birkhäuser; 2015.
18. Angeli D, Sontag ED. Forward completeness, unboundedness observability, and their Lyapunov characterizations. Syst Control Lett.
1999;38:209-217.
19. Levant A. Higher-order sliding modes, differentiation and output-feedback control. Int J Control. 2003;76:924-941.
FERREIRA DE LOZA ET AL. 17

20. Iriarte R, Aguilar LT. Output feedback second-order sliding-mode tracking control for perturbed inertia wheel pendulum.
In: Boubaker O, Iriarte R, eds. The Inverted Pendulum: From Theory to New Innovations in Control and Robotics. London, UK: IET;
2017:137-150.
21. Aguilar L, Boiko I, Fridman L, Freidovich L. Generating oscillations in inertia wheel pendulum via two-relay controller. Int J Robust
Nonlinear Control. 2012;22(3):318-330.
22. Fridman L, Shtessel Y, Edwards C, Yan XG. Higher-order sliding-mode observer for state estimation and input reconstruction in nonlinear
systems. Int J Robust Nonlinear Control. 2008;18:399-412.
23. Rios H, Davila J, Fridman L, Edwards C. Fault detection and isolation for nonlinear systems via high-order-sliding-mode
multiple-observer. Int J Robust Nonlinear Control. 2015;25(16):2871-2893.
24. Kamachkin AM, Potapov DK, Yevstafyeva VV. Existence of periodic solutions to automatic control system with relay nonlinearity and
sinusoidal external influence. Int J Robust Nonlinear Control. 2017;27(2):204-211.
25. Block D, Aström K, Spong M. The Reaction Wheel Pendulum. Synthesis lectures on control and mechatronics 1. San Rafael, CA: Morgan
& Claypool Publishers; 2007.
26. Boiko I. Discontinuous Control Systems: Frequency-Domain Analysis and Design. Boston, MA: Birkhäuser; 2009.

How to cite this article: Ferreira de Loza A, Fridman L, Aguilar LT, Iriarte R. High-order
sliding-mode observer–based input-output linearization. Int J Robust Nonlinear Control. 2019;1–17.
https://doi.org/10.1002/rnc.4556

Вам также может понравиться