Anda di halaman 1dari 22

Available online at www.sciencedirect.

com

Journal of the Franklin Institute 354 (2017) 5328–5349


www.elsevier.com/locate/jfranklin

Output feedback adaptive robust control of hydraulic


actuator with friction and model uncertainty
compensation
Xudong Li, Jianyong Yao∗, Changsheng Zhou
School of Mechanical Engineering, Nanjing University of Science and Technology, Nanjing 210094, PR China
Received 26 January 2017; received in revised form 19 June 2017; accepted 26 June 2017
Available online 3 July 2017

Abstract
This paper studies output feedback control of hydraulic actuators with modified continuous LuGre
model based friction compensation and model uncertainty compensation. An output feedback adaptive
robust controller is constructed which combines the adaptive control part and the robust control part
seamlessly. The adaptive part is constructed to handle the parametric uncertainties existed in the model.
The residuals coming from parameter adaption and the unmodeled dynamics are taken into consideration
by the robust part. Since only the position signal is available, the velocity, pressure, and internal friction
states are obtained by observation or estimation. The errors coming from observation and estimation
are also dealt with the robust part. Furthermore, the convergence of the closed-loop controller–observer
scheme is achieved by the Lyapunov method in the presence of parametric uncertainties only. Extensive
comparative experiments performed on a hydraulic actuator demonstrate the effectiveness of the proposed
controller–observer scheme.
© 2017 The Franklin Institute. Published by Elsevier Ltd. All rights reserved.

1. Introduction

Electro-hydraulic actuators and electro-mechanical actuators have been widely used in


industries [1–3]. With the development of the motor technology, the electro-mechanical actu-

∗ Corresponding author.
E-mail addresses: jerryyao.buaa@gmail.com, yaojianyong1984@163.com (J. Yao).

http://dx.doi.org/10.1016/j.jfranklin.2017.06.020
0016-0032/© 2017 The Franklin Institute. Published by Elsevier Ltd. All rights reserved.
X. Li et al. / Journal of the Franklin Institute 354 (2017) 5328–5349 5329

Nomenclature

Ps , Pr the supply and return pressure


P1 , P2 the pressures in the two chambers
Q1 , Q2 the supply and return flow rate
M the mass of the load
y, y˙, ÿ the displacement, velocity and acceleration of the
load
PL the load pressure
Dm the ram area of the cylinder
Fe the external disturbance
Ff the friction term
σ 0 , σ 1, σ 2 the friction force parameter
z the internal friction state
fs the Coulomb friction
fc the stiction friction
b1 , b2 , b3 the shape coefficients
c1 , c2 the known constants
Vt the total volume of the actuator
βe the effective oil bulk modulus
Ct the coefficient of the total internal leakage
d˜p the modeling error
QL the load flow
kt the total flow gain
u the control input
θ the unknown parameters
θˆ, θ˜ the parameter estimation and estimation errors
x the state variables
xˆ,x˜ the state estimation and estimation errors
h1 , h2 the known constants
 the diagonal matrix
τ the adaption function
η1 , η2 the learning functions
p the auxiliary variable
ω0 , ω1 the control gains
zˆ1 , zˆ2 the friction state estimation
x1d the desired position command
k1 , k2 , k3 , ks2 , ks3 , k4 , k2s , k3s , k4s the positive feedback gains
α2 the virtual control
umax the maximum hardware constraint for the control in-
put
ε1 , ε2 , ε3 the positive design parameters
δ1, δ2, δ3 the unknown but bounded functions

ators have been more popular than electro-hydraulic actuators [4–6]. But the electro-hydraulic
actuators cannot be replaced by the electro-mechanical actuators in some applications
5330 X. Li et al. / Journal of the Franklin Institute 354 (2017) 5328–5349

due to the advantages of the small size-to-power ratios and the ability to sustain large
forces and torques [7], such as heavy engineering machineries [8], aircraft actuators [9],
hardware-in-the-loop simulators [10,11], and so on. So it is necessary to concern about
the high performance control techniques for hydraulic systems. However, the modeling
uncertainties and parameter variations of the hydraulic systems and the nonlinearity exists
in the external environment such as uncompensated friction complicate the development of
high performance of hydraulic control techniques. Lots of advanced control methods have
been proposed to handle various difficulties to achieve high performance, such as feedback
linearization control, adaptive control, robust sliding mode control, adaptive robust control
[12–14]. These nonlinear control methods not only solved the control problem coming from
nonlinear hydraulic systems successfully but also demonstrated that nonlinear controllers can
achieve better performance than conventional linear controllers [7,15,16].
Friction compensation has been the subject of precision servo control in many recent
studies [17–22]. In spite of not all the friction compensation methods need an exact friction
model [23], but the model based friction compensation technique is more popular and has
received greater attention. Among these nonlinear friction models, the LuGre model can
capture the major features of the friction nonlinear dynamics such as the Stribeck effect
and hysteresis [24], and the LuGre model has been widely used for dynamic friction com-
pensation. Some modifications were made for LuGre model for special matters. A modified
LuGre model was proposed by [25] to substitute the original model by a static model at high
speed for solving the digital implementation problems. In order to adapt to the backstepping
controller designing process [26–28], some novel continuously differentiable modified LuGre
models were proposed by [18,29].
However, almost all of the aforementioned nonlinear control techniques and friction
compensation methods need full-state feedback. That is to say, the hydraulic actuator system
should be equipped with position sensor, velocity sensor and pressure sensor in order to
get all states’ information. But for many applications, because of volume/weight limitations
and/or structure restriction, velocity sensor or pressure sensor is not commonly equipped.
And the internal LuGre friction states cannot be measured. So the output controller design
for hydraulic actuator motion control is an imperious demand.
As far as I know, few literatures concerned about the output feedback control of hydraulic
system because of the high nonlinearity existed in hydraulic systems. In this paper, the
unique feature is that output feedback control is implemented combining with modified
LuGre model based friction compensation.
An output feedback adaptive controller with modified continuous LuGre model based
friction compensation and model uncertainty compensation is proposed in this paper. Only
position signal is needed. The velocity state is estimated through a discontinuous observer,
and the other state concerning about the pressure is estimated through the controller–observer
designing process later. A dual observer proposed by [30] is used for the estimation of the
internal friction states. The adaptive robust control with backstepping scheme combines the
adaptive control part and the robust control part seamlessly. The adaptive part is constructed
to handle the parametric uncertainties existed in the model. Then the residuals coming from
parameter adaption and the unmodeled dynamics are taken into consideration by the robust
part. Moreover, the errors coming from state observation or estimation are dealt with the
robust part. Furthermore, the convergence of the closed-loop controller–observer scheme is
achieved by the Lyapunov method in the presence of parametric uncertainties only. Extensive
X. Li et al. / Journal of the Franklin Institute 354 (2017) 5328–5349 5331

Fig. 1. Architecture diagram of the hydraulic system.

comparative experiments performed on a hydraulic actuator demonstrate the effectiveness of


the proposed controller–observer scheme.
The paper is organized as follows. Section 2 gives problem formulation and dynamic
models. Section 3 gives the adaptive robust position controller designing process and the-
oretical results. Section 4 carries experimental setup and comparative experimental results,
and Section 5 contains the conclusions.

2. Problem formulation and dynamic models

The hydraulic actuator system under study is shown in Fig. 1. Ps is the supply pressure
which is constant. Pr is the return pressure. P1 and P2 are the pressures in the two chambers.
Q1 is the supplied flow rate of the forward chamber and Q2 is the return flow rate of the
return chamber.
The dynamics equation of the mass load can be expressed as

M ÿ = PL Dm − Ff + Fe (1)

where M represents the mass of the load; y, y˙ and ÿ represent the displacement, velocity and
acceleration of the load; PL = P1 – P2 represents the load pressure of the hydraulic cylinder;
Dm denotes the ram area of the cylinder. Fe represents external disturbance; Ff represents
the friction term. There have been many friction models proposed in the literature, but they
cannot be used in backstepping controller designing process because of the discontinuous
property. A novel continuously differentiable nonlinear friction model proposed by [18] is
used in this paper, which is given by

Ff = σ0 z + σ1 z˙ + σ2 y˙
z˙ = y˙ − N (y˙)z (2)

where σ 0 , σ 1 and σ 2 represent the friction force parameters which can be physically
interpreted as the stiffness coefficient, damping coefficient and the viscous coefficient; z
represents the unmeasured internal friction state. And the function N (y˙) is defined as

N (y˙) =
g(y˙)
5332 X. Li et al. / Journal of the Franklin Institute 354 (2017) 5328–5349

400

300

200
Static Friction force(Nm)

100

Experimental Data
0 Identification Curve

-100

-200

-300

-400
-0.08 -0.06 -0.04 -0.02 0 0.02 0.04 0.06 0.08
Velocity(m/s)

Fig. 2. The static friction force and its curve fitting.

g(y˙) = ( fs − fc )[tanh (b1 y˙) − tanh (b2 y˙)] + fc tanh (b3 y˙) (3)

where fc represents the Coulomb friction and fs represents the stiction friction; b1 , b2 and b3
represent shape coefficients to approach various friction effects. As mentioned in [18], the
term tanh (b1 y˙) − tanh (b2 y˙) captures the Stribeck effect and the term tanh (b3 y˙) represents the
Coulomb friction coefficient. The coefficients b1 = 500, b2 = 80, b3 = 900, fs = 300, fc = 100
are used to fit the experimental data in Fig. 2.
The load pressure dynamics can be expressed as [2]
Vt
P˙L = −Dm y˙ − Ct PL + QL + d˜p (4)
4βe
where Vt is the total control volume of the actuator; β e is the effective oil bulk modulus; Ct
is the coefficient of the total internal leakage of the actuator due to pressure; d˜p is modeling
error; QL is the load flow expressed as.

QL = kt u Ps − sign(u)PL (5)

where kt is the total flow gain with respect to u.


The state variables of the hydraulic actuator system are chosen as x = [x1 , x2 , x3 ]T =
[y, y˙, Dm PL /M]T . Then the state-space form of the entire system can be expressed as

x˙1 = x2
σ0 σ1 σ1 + σ2 Fe
x˙2 = x3 − z + N ( x2 )z − x2 +
M M M M
X. Li et al. / Journal of the Franklin Institute 354 (2017) 5328–5349 5333

4Dm βe kt M 4Dm2 βe 4βe 4Dm βe ˜
x˙3 = Ps − sign(u) x3 u − x2 − Ct x3 + dp (6)
MVt Dm MVt Vt MVt

3. Adaptive robust control of hydraulic actuator systems

3.1. Design model and issues to be addressed

In order to design the controller more conveniently, the system model can be reformulated
as follow:
Define the unknown parameter set θ = [θ 1 ,θ 2 ,θ 3 ]T , where θ 1 = σ 0 /M, θ 2 = σ 1 /M,
θ 3 = (σ 0 + σ 1 )/M. It should be noted that other parameters are all known. Thus, the
state-space equation can be formulated as:

x˙1 = x2
x˙2 = x3 − θ1 z + θ2 N (x2 )z − θ3 x2 + d1
x˙3 = g(u, x3 )u − c1 x2 − c2 x3 + d2 (7)

where
4Dm2 βe 4βe Fe 4Dm βe ˜
c1 = c2 = Ct d1 = d2 = dp
MVt Vt M MVt

4Dm βe kt M
g(u, x3 ) = Ps − sign(u) x3 (8)
MVt Dm

Assumption 1. The parameter set defined above and modeling uncertainties satisfy

θ ∈ θ {θ : θmin ≤ θ ≤ θmax }⎬
|d1 | ≤ δf1 (9)

|d2 | ≤ δf2

where θ min = [θ 1min ,…, θ 3min ]T , θ max = [θ 1max ,…, θ 3max ]T ; δ f1 and δ f2 are some positive
constants.
Remark 1. Define ψ (x2 ) = θ2 N (x2 )z − θ3 x2 . It is easy to know that the function ψ(x2 ) is
Lipschitz with respect to x2 . And g(u,x3 ) is Lipschitz with respect to x3 in its practical range.
x˜2 and x˜3 represent the estimation error which are defined in the section below. There exist
two known constants h1 , h2 such that:
 
 ˜
ψ  ≤ h1 |x˜2 | |g˜| ≤ h2 |x˜3 | (10)

where ψ˜ = ψ (xˆ2 ) − ψ (x2 ); g˜ = g(u, xˆ3 ) − g(u, x3 ).

3.2. Projection mapping and parameter adaptation

The parametric uncertainty is an important issue for model-based control. Let θˆ represent
the estimate of θ and θ˜ represent the estimation error (i.e. θ˜ = θˆ − θ ). Combining Eq. (9),
a projection mapping can be defined as [15]
5334 X. Li et al. / Journal of the Franklin Institute 354 (2017) 5328–5349

⎨0 if θˆi = θi max and •i > 0
Projθˆi (•i ) = 0 if θˆi = θi min and •i < 0 (11)

•i otherwise
where i = 1,…,3. In this paper, a parameter adaptation law is given by
˙
θˆ = Projθˆ (τ ), θmin ≤ θˆ (0) ≤ θmax (12)
with
Projθˆ (•) = [Projθˆ (•1 ), ..., Projθˆ (•3 )]T (13)
where Г > 0 is a diagonal matrix, and τ is an adaptation function to be designed later. For
any adaption function τ , the projection mapping expressed in Eq. (12) guarantees

(P1 ) θˆ ∈ θˆ θˆ : θmin ≤ θˆ ≤ θmax
(14)
(P2 ) θ˜T [ −1 Proj ˆ (τ ) − τ ] ≤ 0,∀τ
θ
Remark 2. The projective estimation has been mentioned in previous literatures [7,30–31].
Property P1 in Eq. (14) means that the parameter estimation stays in a known bounded region
which is very important for the robust control law designed below. In addition, it is also a
protective measure for adaptive control. Since large disturbance might cause the adaptive law
to be unstable if no ad hoc protection solution. Property P2 in Eq. (14) shows that the use of
projection modification to the traditional discontinuous adaption law holds the perfect learning
capability of the traditional one.

3.3. State estimation

As mentioned at the beginning of the paper, only the position signal is available. That
is to say the velocity state, pressure state, internal friction state should be obtained by the
synthesized observer or estimator. The velocity state x2 is estimated by the discontinuous
velocity observer described in [32]. The first two equations of Eq. (9) are utilized to design
the discontinuous velocity observer which provides exponentially convergent estimates of the
velocity state x2 . Let xˆ2 represent the estimate of x2 and x˜2 represent the estimation error
(i.e. x˜2 = xˆ2 − x2 ). Let xˆ1 represent the estimate of x1 and x˜1 represent the estimation of
error (i.e. x˜1 = xˆ1 − x1 ). The discontinuous velocity observer has the following structure:
xˆ2 = x˙ˆ1 = p + ω0 x˜1
p˙ = ω1 sgn (x˜1 ) + (ω0 − 1)x˜1 (15)
where p is an auxiliary variable; ω0 , ω1 are constant. The estimation of pressure state x3 will
be obtained in the following sections.
Remark 3. The global asymptotic regulation of x˜2 will be guaranteed by the velocity observer
described by Eq. (15) which means that x˜2 → 0 as t → ∞ [32]. That is to say, a bounded
velocity state estimation error can be obtained for any time t > 0. So in the controller designing
process below, x˜2 will be treated as a bounded variable.
As with the velocity or pressure state, the friction state z is also unknown. The internal
unmeasured state z is related two different terms. So in order to estimate the state of z, a
robust dual-observer with a projection type is used which was proposed in [30].
z˙ˆ1 = projzˆ (η1 ), z˙ˆ2 = projzˆ (η2 )
1 2
(16)
X. Li et al. / Journal of the Franklin Institute 354 (2017) 5328–5349 5335

where zˆ1 , zˆ2 are the estimations of the internal friction states z1 and z2 ; η1 , η2 are the learning
functions which will be synthesized later; the projection is similar to the projection used in
the section before which also has the similar properties described in (14).

⎨0 if zˆi = zmax and ηi > 0
Projzˆi (ηi ) = 0 if zˆi = zmin and ηi < 0 (17)

ηi otherwise
According to [30] the physical bounds of the friction state z can be set as zmax = fs ,
zmin = −fs .

3.4. Controller design

The design parallels the backstepping design because of the unmatched uncertainty in
Eq. (7).
Step 1: Define a switching-function-like quantity as:

e2 = e˙1 + k1 e1 = x2 − x2eq , x2eq = x˙1d − k1 e1 (18)

where e1 = x1 − x1d is the output tracking error; x1d is the desired position command; k1 is a
positive feedback gain. The dynamic of e2 can be given as

e˙2 = x˙2 − x˙2eq = x3 − θ1 z + θ2 N (xˆ2 )z − θ3 xˆ2 − ψ˜ + d1 − ẍ1d + k1 x˙1 − k1 x˙1d (19)

Let e3 = x3 − α 2 denote the input discrepancy, then Eq. (19) can be rewritten as

e˙2 = e3 + α2 − θ1 z + θ2 N (xˆ2 )z − θ3 xˆ2 − ψ˜ + d1 − ẍ1d + k1 x˙1 − k1 x˙1d (20)

In this step, x3 can be treated as a virtual control input and the virtual control law α 2 can
be designed for it which consists of two terms given by

α2 = α2a + α2s
α2a = θˆ1 zˆ1 − θˆ2 N (xˆ2 )zˆ2 + θˆ3 xˆ2 + ẍ1d − k1 xˆ2 + k1 x˙1d
α2s = α2s1 + α2s2
α2s1 = −k2 (xˆ2 − x2eq ) (21)

where k2 > 0 is a feedback gain.


In Eq. (21), α 2a acts as a model-based adaptive control law. α 2s acts as a robust control law
which contains a linear part and a nonlinear part. Substituting Eq. (21) into Eq. (20), we have

e˙2 = e3 + α2s2 − k2 e2 + θ˜T ϕ1 + θ1 z˜1 − θ2 N (x2 )z˜2 − (k2 + k1 )x˜2 − ψ˜ + d1 (22)

ϕ1 = [zˆ1 , −N (xˆ2 )zˆ2 , xˆ2 ]T (23)

where z˜1 , z˜2 are the estimation errors of the internal friction states (i.e. z˜1 = zˆ1 − z1 ,
z˜2 = zˆ2 − z2 ).
Define a positive semi-definite (p.s.d.) function V1 as
1 2 1 2
V1 = e + e (24)
2 1 2 2
5336 X. Li et al. / Journal of the Franklin Institute 354 (2017) 5328–5349

From Eq. (24), its time derivative satisfies

V˙1 = e2 e3 − k1 e21 + e1 e2 − k2 e22 + e2 (α2s2 + θ˜T ϕ1 − (k2 + k1 )x˜2 + θ1 z˜1 − θ2 N (xˆ2 )z˜2 − ψ˜ + d1 )
(25)

Thus there exists a robust control function satisfying the following condition:

e2 (α2s2 + θ˜T ϕ1 − (k2 + k1 )x˜2 + θ1 z˜1 − θ2 N (xˆ2 )z˜2 − ψ˜ + d1 ) ≤ ε1 δ12 (26)

where ε1 is a positive design parameter which can be arbitrarily small, and δ 1 is an unknown
but bounded function satisfies:
 
 1 

δ1 ≥ θM ϕ1 + k2 + k1 + |x˜2 | + h1 |x˜2 | + θ1M zM + θ2M N2 (xˆ2 )zM + δf1 (27)
4 ε1 
in which θ M = θ max − θ min , θ 1M = θ 1max − θ 1min , θ 2M = θ 2max − θ 2min , zM = zmax − zmin .
Then α 2s2 can be chosen as
1
α2s2 = −ks2 (xˆ2 − x2eq ) = − (xˆ2 − x2eq ) (28)
4 ε1
Step 2: This is the final design step in which the actual control u will be synthesized such
that x3 tracks the desired virtual control function α 2 . The derivative of e3 can be obtained as:

e˙3 = x˙3 − α˙ 2 = g(u, x3 )u + ϕ(x2 , x3 ) + d2 − α˙ 2c − α˙ 2u


= g(u, xˆ3 )u + ϕ(xˆ2 , xˆ3 ) + d2 − α˙ 2c − α˙ 2u − g˜u + c1 x˜2 + c2 x˜3 (29)

where xˆ3 is the estimation of x3 and x˜3 represents the estimation error (i.e. x˜3 = xˆ3 − x3 )
∂ α2 ∂ α2 ˆ˙ ∂ α2 ˙ ∂ α2 ˙ ∂ α2 ∂ α2 ˙
α˙ 2c = + θ+ zˆ1 + zˆ2 + xˆ2 + xˆ2
∂t ∂ θˆ ∂ zˆ1 ∂ zˆ2 ∂ x1 ∂ xˆ2
∂ α2
α˙ 2u = x˜2 (30)
∂ x1
Similar to above designing process, we can derive an actual control input u:

u = ua + us
1
ua = [−ϕ(xˆ2 , xˆ3 ) + α˙ 2c ]
g(u, xˆ3 )
1
us = (us1 + us2 ), us1 = −k3 (xˆ3 − α2 ) (31)
g(u, xˆ3 )
where k3 > 0 is a feedback gain.
As mentioned before, ua is utilized for model compensation, us1 and us2 are designed to
stabilize the system. Substituting Eq. (31) into Eq. (29), the dynamic of e3 changes to

e˙3 = us2 − k3 e3 − k3 x˜3 − α˙ 2u + d2 − g˜u + c1 x˜2 + c2 x˜3 (32)

Consider an augmented p.s.d. function V2 given by


1 1
V2 = V1 + e23 + x˜32 (33)
2 2
X. Li et al. / Journal of the Franklin Institute 354 (2017) 5328–5349 5337

The time derivative of V2 can be given by


V˙2 = V˙1 − k3 e23 + e3 (us2 − k3 x˜3 − α˙ 2u + d2 − g˜u + c1 x˜2 + c2 x˜3 )
+ x˜3 (x˙ˆ3 − g(u, xˆ3 )u − ϕ(xˆ2 , xˆ3 ) − d2 + g˜u − c1 x˜2 − c2 x˜3 ) (34)
Similar to Eq. (27), the robust control function us2 ˙
and the state estimation xˆ3 are chosen
to satisfy
Condition i:
 
1
e3 us2 − α˙ 2u + d2 + x˜3 + c1 x˜2 + c2 x˜3 ≤ ε2 δ22 (35)
4 ε2
Condition ii:
 
1
x˜3 (x˙ˆ3 − g(u, xˆ3 )u − ϕ(xˆ2 , xˆ3 ) − d2 + + h2 umax e3 + g˜u − c1 x˜2 ) ≤ ε3 δ32 (36)
4 ε3
where ε 2 , ε 3 are positive design parameters which can be arbitrarily small; umax is a positive
constant which presents the maximum hardware constraint for the control input; δ 2 and δ 3
are unknown but bounded functions satisfying:
  
 ∂ α2 
δ2 ≥   + c1 |x˜2 | + δf2 δ3 ≥ c1 |x˜2 | + δf2 (37)
∂x 1

Then us2 and x˙ˆ3 are given by


x˙ˆ3 = g(u, xˆ3 )u + ϕ(xˆ2 , xˆ3 ) + x3s
 
1
x3s = −(k4 + h2 umax )(xˆ3 − α2 ) = − + h2 umax (xˆ3 − α2 ) (38)
4 ε3
1
us2 = −(ks3 + c2 )(xˆ3 − α2 ) ks3 = (39)
4 ε2
For clear presentation, the block diagram of the proposed controller–observer scheme is
summarized in Fig. 3 as follow.

3.5. Main result


Theorem. Let the parameter estimates be updated by the projection type adaptation law
(Eq. (12)) for parametric uncertainties. The adaptation function of τ is chosen as
τ = ϕ1 (x2eq − xˆ2 ) (40)
The projection type state observation and learning functions are given by
η1 = xˆ2 − N (xˆ2 )zˆ1 − r1 (xˆ2 − x2eq ) (41)

η2 = xˆ2 − N (xˆ2 )zˆ1 + r2 (xˆ2 − x2eq )N (xˆ2 ) (42)


where r1 , r2 are learning gains, and choosing feedback gains k1 , k2 , k3 , k4 such that the
matrix below is positive definite; where K = k3 + ks3 + k4 + 2h2 umax
⎡ ⎤
k1 − 21 0 0
⎢− 1 k2 − 21 0 ⎥
=⎢ ⎣ 0
2 ⎥ (43)
−2 1
k3 − K2 ⎦
0 0 − K2 k4
5338 X. Li et al. / Journal of the Franklin Institute 354 (2017) 5328–5349

Fig. 3. Block diagram of the proposed controller–observer.

A. Under normal circumstances, the modeling errors always exist which is to say that d1
= 0,
d2
= 0. Then the proposed control law guarantees that the control input and all internal
signals are bounded. Furthermore, V2 is bounded above by

ε1 ||δ1 ||2∞ + ε2 ||δ2 ||2∞ + ε3 ||δ3 ||2∞


V2 (t ) ≤ exp(−κt )V2 (0) + [1 − exp(−κt )] (44)
κ
where κ = −2λmin (), ||δ 1 ||∞ ,||δ 2 ||∞ , and ||δ 3 ||∞ stand for the infinity norm of δ 1 (t), δ 2 (t)
and δ 3 (t).

B. If after a finite time t0 , in the presence of parametric uncertainties only (i.e., d1 =


0, d2 = 0, x˜2 = 0, ∀t ≥ t0 ), then, asymptotic output tracking is also achieved (i.e. e → 0
as t → ∞), where e = [‫׀‬e1 ‫׀‬,‫׀‬e2 ‫׀‬,‫׀‬e3 ,‫|׀‬x˜3 |]T .

Proof of the theorem is given in the appendix


Remark 4. Two results corresponding to two different situations can be achieved by the
control scheme proposed in this paper.
Under normal circumstances, the modeling errors always exist. Then the results in A of the
Theorem indicate that the boundedness of the tracking error can be guaranteed. The transient
performance and the final tracking error can be adjusted via adjusting certain controller param-
eters. In fact, even if the adaptive mechanism does not work, the results in A of the Theorem
are still valid. So it is a well-designed robust controller by the results in A of the Theorem.
X. Li et al. / Journal of the Franklin Institute 354 (2017) 5328–5349 5339

Fig. 4. Experimental platform of double-rod hydraulic actuator system.

When the model uncertainties and the velocity estimation error does not exist (i.e.,
d1 = 0, d2 = 0, x˜2 = 0,∀t ≥ t0 ). The results in B of the Theorem show that the asymptotic
tracking performance can be achieved which is contributed to the parameter adaptive mech-
anism which eliminates the parametric uncertainties. So it is also a well-designed adaptive
controller by the results in B of the Theorem.
The results fully illustrate that the robust control scheme and the adaptive control scheme
in this paper is combined seamlessly.

4. Comparative experimental results

4.1. Experimental setup

In this section, a verification platform equipped with a double-rod hydraulic actuator


is shown in Fig. 4. This platform consists of a bench case, a hydraulic cylinder, a linear
encoder and a high bandwidth servo valve. The specification of the hardware components
can be found in Table 1. The monitor software and the real time control software make up
the measurement and control system. The sampling time τ is 0.5 ms.

4.2. Comparative experimental results

The following three controllers are compared.


OFALuGre: This is the output feedback adaptive robust controller with LuGre
model based friction compensation (Eq. (31)) proposed in this paper and described in
Section 3. For simplicity, the nonlinear control functions α2s = −(k2 + ks2 )(xˆ2 − x2eq ),
us = − g(u,1xˆ3 ) (k3 + ks3 + c2 )(xˆ3 − α2 ) and x3s = −(k4 + h2 umax )(xˆ3 − α2 ) are implemented as
5340 X. Li et al. / Journal of the Franklin Institute 354 (2017) 5328–5349

Table 1
Specification of the double-rod hydraulic actuator system.

Component Specification
Hydraulic supply Supply pressure 100 bar
Servo valve Type Moog G761-3003
Rated flow 19 L/min at 70 bar drop
Bandwidth ≥120 Hz
Hydraulic actuator Stroke ±44 mm
Efficient ram area 904.78 mm2
Load mass Mass 30 kg
Linear encoder Type Heidenhain LC483
Accuracy ±5 μm
A/D card Type Advantech PCI-1716
D/A card Type Advantech PCI-1723
Counter card Type Heidenhain IK-220
Computer Type IEI WS-855GS

α2s = −k2s (xˆ2 − x2eq ), us = −k3s (xˆ3 − α2 ) and x3s = −k4s (xˆ3 − α2 ). The feedback gains are
chosen large enough which is normal for practical application. The control gains are given as:
k1 = 500, k2s = 3000, k3s = 0.015, k4s = 400, ω0 = 300, ω1 = 0.01. The boundary of uncertain
ranges are given by θ max =[107 , 1500, 200]T , θ min = [−107 , 500, 100]T . The initial estimate
of θ is chosen as θˆ (0) =[8 × 106 , 1200, 150]T . Adaptation rates are set as Г = diag{2 × 107 ,
4 × 108 , 2 × 103 }. Friction state estimation rates are set as: r1 = 2 × 10−4 , r2 = 2 × 10−4 . The
bounds of z estimation are set as: zmax = 2 × 10−5 , zmin = −2 × 10−5 .
OFAC: This is the output feedback controller without friction compensation compared
with OFALuGre controller (i.e. zˆ1 = zˆ2 = 0). The controller parameters are the same as the
corresponding parameters in OFALuGre controller.
ESOC: This is the ESO based output feedback controller which was proposed by [33].
The state variables are defined as x = [x1 , x2 , x3 ]T = [y, y˙, Dm PL /M]T . The structure of the
controller is shown below
 
1 4βe (Dm 2 xˆ2 + Ct M xˆ3 )
u= − xˆ4 + α˙ 2c − v3 (xˆ3 − α2 )
g(u, xˆ3 ) MVt
4000xˆ2 + 100atan(900x˙1d )
α2 = + ẍ1d − v1 xˆ2 + v1 x˙1d − v2 (xˆ2 − x2eq )
M
∂ α2 ∂ α2 ∂ α2 ˙
α˙ 2c = + xˆ2 + xˆ2
∂t ∂ x1 ∂ xˆ2
x2eq = x˙1d − v1 z1 z1 = x1 − x1d (45)
and x2 , x3 are estimated by an extended state observer defined as
x˙ˆ = A0 xˆ + (xˆ ) + Gu + (x1 − xˆ1 ) (46)
where
 T
4000xˆ2 + 100atan(900x˙1d )
(xˆ ) = 0, − , −c1 xˆ2 − c2 xˆ3 , 0
m
G = [0, 0, g(u, xˆ3 ), 0]T
H = [4ω0 , 6ω02 , 4ω03 , ω04 ]T (47)
X. Li et al. / Journal of the Franklin Institute 354 (2017) 5328–5349 5341

Table 2
Performance indexes during the last cycle for normal motion.

Controller Me μ σ
OFALuGre 0.0664 0.0338 0.0134
OFAC 0.0969 0.0453 0.0150
ESOC 0.1553 0.0599 0.0388
PI 0.2628 0.1294 0.0815

The controller gains are given as: v1 = 500, v2 = 400, v3 = 200, ω0 = 200.
PI: This is the proportional-integral controller. The controller gains tuned carefully via
error-and-try method are kp = 1500, ki = 500.

Remark. 5. Some simplifications are made to select the special robust control gains in the
controller implementation. We may implement the needed robust gains in two ways. The first
way is to choose the robust gains rigorously to ensure the theoretical stringency with various
prerequisites. However, it increases the complexity of the resulting control law. In addition
to that, a simple but effective way is used in this paper which simply chooses these control
gains large enough without worrying about the specific prerequisites. The results in A of the
theorem indicate that the bigger the control gains are, the smaller the final tracking error is.
In this way, prerequisites (Eqs. (27), (35) and (36)) will be satisfied for certain set of values
of k1 , k2 , k3 , k4 , ks2 , ks3 . With this simplification strategy, the resulting OFALuGre controller
is yet sufficient as shown in the comparative experimental results.

When we conduct the experiments, the PID controller is implemented on the experimental
platform first. While the PID controller is running, the proposed OFALuGre controller is
running on the side using the position error signal which is produced by the PID controller.
Through the visualization window, we can get the control output of the OFALuGre controller.
By adjusting the feedback gains of the OFALuGre controller, a result that matches the reality
can be achieved. Then the OFALuGre controller can be implemented on the experimental
platform.
Some indexes are used to assess the quality of the control algorithm: the maximal value
of tracking error Me, the average value of tracking error μ and the standard deviation value
of the tracking error σ which are defined in [2].
The four controllers are first tested for a smooth normal motion trajectory
x1d (t) = 10arctan(sin3.14t)•[1 − exp(−t)]/(π /4). The control performance of OFALuGre is
shown in Fig. 5. The corresponding tracking performances of the compared controllers are
presented in Fig. 6. The performance indexes of four controllers during the last cycle are
presented in Table 2. Obviously, these three nonlinear controllers perform better than PI
controller. All results demonstrate that the OFALuGre controller handles better than the other
three controllers. When comparing the performance of OFALuGre with OFAC, it can be seen
that OFALuGre is better than OFAC because of the friction compensation. It can be indicated
that the LuGre model based friction compensation effectively suppresses the effects of the
friction. The control input of all controllers are presented in Fig. 7. And the convergence of
the parameter estimation of OFALuGre controller is shown in Fig. 8. When compared with
ESOC, OFALuGre and OFAC controller show better performance which demonstrate that
the output feedback control scheme handles better than ESO scheme.
5342 X. Li et al. / Journal of the Franklin Institute 354 (2017) 5328–5349

10
x
1d
5 x
Position(mm)

-5

-10
0 5 10 15 20 25 30 35 40 45 50
time(s)

0.1
OFALuGre Error(mm)

0.05

-0.05

-0.1
0 5 10 15 20 25 30 35 40 45 50
time(s)

Fig. 5. Tracking performance of OFALuGre for normal motion.


OFAC Error(mm)

0.2

-0.2
0 5 10 15 20 25 30 35 40 45 50
time(s)
ESOC Error(mm)

0.2

-0.2
0 5 10 15 20 25 30 35 40 45 50
time(s)
0.5
PI Error(mm)

-0.5
0 5 10 15 20 25 30 35 40 45 50
time(s)

Fig. 6. Tracking errors of other three controllers for normal motion.


OFALuGre Control Input(V) X. Li et al. / Journal of the Franklin Institute 354 (2017) 5328–5349 5343

2 2

OFAC Control Input(V)


1 1

0 0

-1 -1

-2 -2
0 10 20 30 40 50 0 10 20 30 40 50
time(s) time(s)

2 1
ESOC Control Input(V)

PI Control Input(V)
0.5
1
0
0
-0.5
-1
-1

-2 -1.5
0 10 20 30 40 50 0 10 20 30 40 50
time(s) time(s)
Fig. 7. Control inputs of all controllers for normal motion.

Fig. 8. Parameter estimation of ARC controller for normal motion.


5344 X. Li et al. / Journal of the Franklin Institute 354 (2017) 5328–5349

10
x
1d
5
Position(mm)

x
1

-5

-10
0 5 10 15 20 25 30 35 40 45 50
time(s)

0.1
OFALuGre Error(mm)

0.05

-0.05

-0.1
0 5 10 15 20 25 30 35 40 45 50
time(s)

Fig. 9. Tracking performance of OFALuGre for slow motion.

Table 3
Performance indexes during the last cycle for slow motion.

Controller Me M σ
OFALuGre 0.0380 0.0215 0.0130
OFAC 0.0722 0.0515 0.0172
ESOC 0.0728 0.0408 0.0153
PI 0.1104 0.0473 0.0336

As we all know, when the motion velocity is slow, the friction nonlinearity
is the main factor influencing the tracking performance. A slow motion trajectory
x1d (t) = 10arctan(sin1.256t)•[1 − exp(−t)](π /4) is used to further test the control scheme
in the proposed algorithm in slow work conditions. The tracking performance of OFALuGre
for slow motion is shown in Fig. 9, whereas that of the other two controllers is shown in
Fig. 10. The control input of all controllers for slow motion are shown in Fig. 11. The
performance indexes are collected in Table 3. As seen, the proposed OFALuGre scheme
shows a perfect performance contributed to the friction compensation. When comparing
OFALuGre controller and OFAC controller, the results demonstrate that the friction compen-
sation mechanism is well-designed and valid. And the output feedback mechanism proposed
in this paper still handles better compared with ESOC controller.

5. Conclusion

An output feedback adaptive robust controller has been proposed for position control of
the hydraulic actuator. Friction compensation issue and output feedback issue are the focus
of this paper. The velocity state is estimated through a discontinuous observer, and the
other state concern about the pressure is estimated through the controller–observer designing
X. Li et al. / Journal of the Franklin Institute 354 (2017) 5328–5349 5345
OFAC Error(mm)

0.1

-0.1
0 5 10 15 20 25 30 35 40 45 50
time(s)
ESOC Error(mm)

0.2

-0.2
0 5 10 15 20 25 30 35 40 45 50
time(s)
0.2
PI Error(mm)

-0.2
0 5 10 15 20 25 30 35 40 45 50
time(s)

Fig. 10. Tracking errors of other three controllers for slow motion.
OFALuGre Control Input(V)

0.6 0.6
OFAC Control Input(V)

0.4 0.4

0.2 0.2

0 0

-0.2 -0.2

-0.4 -0.4
0 10 20 30 40 50 0 10 20 30 40 50
time(s) time(s)

1 0.2
ESOC Control Input(V)

PI Control Input(V)

0
0.5
-0.2

-0.4
0
-0.6

-0.5 -0.8
0 10 20 30 40 50 0 10 20 30 40 50
time(s) time(s)

Fig. 11. Control inputs of all controllers for slow motion.


5346 X. Li et al. / Journal of the Franklin Institute 354 (2017) 5328–5349

process. A dual observer is used for the internal friction states. The adaptive robust controller
with backstepping technique is constructed to take into account not only parameter variations
but also the unmodeled dynamics and the errors coming from the observation or estimation.
Furthermore, the convergence of the closed-loop controller–observer scheme is demonstrated
by the Lyapunov method in the presence of parametric uncertainties only. The effectiveness
of the proposed strategy is illustrated by extensive comparative experiments.

Acknowledgments

This work was supported in part by the National Natural Science Foundation of China
(grant no. 51675279), in part by the Fundamental Research Funds for the Central Universities
(grant no. 30917011205), and in part by the Young Elite Scientist Sponsorship Program by
CAST (grant no. 2016QNRC001).

Appendix

Proof of Theorem:
From Eqs. (18), (22), (32) and (38), it follows that:
V˙2 = e1 (e2 − k1 e1 ) + e2 (e3 + α2s2 − k2 e2 + θ˜T ϕ˜1 + θ1 z˜1 − θ2 N (x2 )z˜2 − (k2 + k1 )x˜2 − ψ˜ + d1 )
+ e3 (us2 − k3 e3 − k3 x˜3 − α˙ 2u + d2 − g˜u + c1 x˜2 + c2 x˜3 )
+ x˜3 (x˙ˆ3 − g(u, xˆ3 )u − ϕ(xˆ2 , xˆ3 ) − d2 + g˜u − c1 x˜2 − c2 x˜3 )
 
1
= −k1 e21 + e1 e2 − k2 e22 + e2 e3 − k3 e23 − k3 x˜3 e3 − c2 + x˜2
4 ε3 3
 
1 1
− + + h2 umax x˜3 e3 − g˜ue3
4 ε2 4 ε3
+ e2 (α2s2 + θ˜T ϕ˜1 + θ1 z˜1 − θ2 N (x2 )z˜2 − (k2 + k1 )x˜2 − ψ˜ + d1 )
1
+ e3 (us2 − α˙ 2u + d2 + x˜3 + c1 x˜2 + c2 x˜3 )
4 ε2
   
˙ 1 1
+ x˜3 xˆ3 − g(u, xˆ3 )u − ϕ(xˆ2 , xˆ3 ) − d2 + x˜3 + + h2 umax e3 + g˜u − c1 x˜2 (48)
4 ε3 2 ε3
And noting the conditions of Eqs. (26), (35) and (36), we can have
  
1 1
V˙2 ≤ −k1 e21 + e1 e2 − k2 e22 + e2 e3 − k3 e23 − k3 + + + h2 umax x˜3 e3
4 ε2 4 ε3
1 2 2
−g˜ue3 − x˜ ε1 δ + ε2 δ22 + ε3 δ32 ≤ −k1 |e1 |2 + |e1 ||e2 | − k2 |e2 |2 + |e2 ||e3 | − k3 |e3 |2
4 ε3 3 1
  
 1 1  1 2

+ k 3 + + + 2h2 umax |x˜3 ||e3 | − x˜ + ε1 δ12 + ε2 δ22 + ε3 δ32
4 ε2 4 ε3 4 ε3 3
⎡ ⎤
k1 − 21 0 0
 ⎢− 1 k2 − 21 0 ⎥  T
= − |e1 ||e2 ||e3 ||x˜3 | ⎢
⎣ 0
2 ⎥
K ⎦ |e1 ||e2 ||e3 ||x˜3 | + ε1 δ12 + ε2 δ22 + ε3 δ32
−2 1
k3 −2
0 0 − K2 k4
≤ −λmin ()(|e1 |2 + |e2 |2 + |e3 |2 + |x˜3 |2 ) + ε1 δ12 + ε2 δ22 + ε3 δ32 (49)
X. Li et al. / Journal of the Franklin Institute 354 (2017) 5328–5349 5347

which leads to Eq. (44). This proves A of Theorem. Now consider the situation in B of
Theorem, i.e. d1 = 0, d2 = 0, x˜2 = 0,∀t ≥ t0 , Define a new Lyapunov function as
1 1 1
V3 = V2 + θ˜T  −1 θ˜ + θ1 r1−1 z˜12 + θ2 r2−1 z˜22 (50)
2 2 2
Then the derivative of V3 becomes
˙
V˙3 = V˙2 + θ˜T  −1 θˆ + r1−1 θ1 z˜1 z˙˜1 + r2−1 θ2 z˜2 z˙˜2
 
1 1
= −k1 e21
+ e1 e2 − k2 e22 + e2 e3 − k3 e23 − k3 x˜3 e3 − c2 x˜32
− x˜3 e3 − + h2 umax x˜3 e3
4 ε2 4 ε3
 
1 2 1
− g˜ue3 − x˜ + e2 − e2 + θ˜T ϕ1 + θ1 z˜1 − θ2 N (x2 )z˜2
4 ε3 3 4 ε1
       
1 4βe 1
+ e3 − + e3 + x˜3 − + h2 umax x˜3 + g˜u
4 ε2 Vt 4 ε3
˙
+ θ˜T  −1 θˆ + r1−1 θ1 z˜1 z˙˜1 + r2−1 θ2 z˜2 z˙˜2
 
1 1
≤ −k1 e21 + e1 e2 − k2 e22 + e2 e3 − k3 e23 − k3 x˜3 e3 − c2 x˜32 − x˜3 e3 − + h2 umax x˜3 e3
4 ε2 4 ε3
1 2 ˙
− g˜ue3 − x˜ +, e2 (θ˜T ϕ1 + θ1 z˜1 − θ2 N (xˆ2 )z˜2 ) − h2 umax x˜32 + g˜ux˜3 + θ˜T  −1 θˆ
4 ε3 3
+ r1−1 θ1 z˜1 z˙˜1 + r2−1 θ2 z˜2 z˙˜2
 
1 1
≤ −k1 e1 + e1 e2 − k2 e2 + e2 e3 − k3 e3 − k3 x˜3 e3 − c2 x˜3 −
2 2 2 2
x˜3 e3 − + h2 umax x˜3 e3
4 ε2 4 ε3
1 2 ˙
− g˜ue3 − x˜ + e2 θ˜T ϕ1 + θ˜T  −1 θˆ + r1−1 θ1 z˜1 z˙˜1 + e2 θ1 z˜1 + r2−1 θ2 z˜2 z˙˜2 − e2 θ2 N (xˆ2 )z˜2
4 ε3 3
(51)
Noting the definition of η1 , η2 , and τ , we can have
˙ ˙
e2 θ˜T ϕ1 + θ˜T  −1 θˆ = θ˜T  −1 (e2 ϕ1 + θˆ ) = θ˜T  −1 (e2 ϕ1 + Projθˆ (ϕ1 (x2eq − xˆ2 )))
= θ˜T  −1 (e2 ϕ1 + Proj ˆ (−e2 ϕ1 ))θ
≤0 (52)

r1−1 θ1 z˜1 z˙˜1 + e2 θ1 z˜1 = r1−1 θ1 z˜1 (z˙ˆ1 − z˙1 ) + e2 θ1 z˜1



= r1−1 θ1 z˜1 z˙ˆ1 − [x2 − N (x2 )(zˆ1 − z˜1 ) − r1 e2 ]

= r1−1 θ1 z˜1 z˙ˆ1 − [x2 − N (x2 )zˆ1 − r1 e2 ] − r1−1 θ1 N (x2 )z˜12
 
= r1−1 θ1 z˜1 projzˆ1 (η1 ) − [x2 − N (xˆ2 )zˆ1 − r1 (xˆ2 − x2eq )] − r1−1 θ1 N (x2 )z˜12
≤ −r1−1 θ1 N (x2 )z˜12 (53)

r2−1 θ2 z˜2 z˙˜2 − e2 θ2 N (xˆ2 )z˜2 = r2−1 θ2 z˜2 (z˙ˆ2 − z˙2 ) − e2 θ2 N (xˆ2 )z˜2

= r2−1 θ2 z˜2 z˙ˆ2 − [x2 − N (x2 )(zˆ2 − z˜2 ) + r2 N (xˆ2 )e2 ]
5348 X. Li et al. / Journal of the Franklin Institute 354 (2017) 5328–5349

= r2−1 θ2 z˜2 zˆ˙2 − [x2 − N (x2 )zˆ2 + r2 N (xˆ2 )e2 ] − r2−1 θ2 N (x2 )z˜22
 
= r2−1 θ2 z˜2 projzˆ2 (η2 ) − [xˆ2 − N (xˆ2 )zˆ2 + r2 N (xˆ2 )(xˆ2 − x2eq )] − r2−1 θ2 N (x2 )z˜22
≤ −r2−1 θ2 N (x2 )z˜22 (54)

Noting that the nonlinear function N(x2 ) is always positive, and as the definition of θ 1 , θ 2
which represent some specific physical meanings, so they are positive too. Then rearrange
Eq. (51), we can have:

V˙3 ≤ −λmin ()(|e1 |2 + |e2 |2 + |e3 |2 + |x˜3 |2 ) − r1−1 θ1 N (x2 )z˜12 − r2−1 θ2 N (x2 )z˜22
= −W ≤ 0 (55)

where λmin () is the minimal eigenvalue of matrix . Therefore, W ∈ L2 and V3 ∈ L∞ . Since
all signals are bounded, it is easy to check that W˙ is bounded and thus uniformly continuous.
So, W → 0 as t → ∞ by Barbalat’s lemma, which leads to B of Theorem.

References

[1] Z. Chen, B. Yao, Q. Wang, Accurate motion control of linear motors with adaptive robust compensation of
nonlinear electromagnetic field effect, IEEE-ASME Trans. Mech 18 (3) (2011) 1122–1129.
[2] J. Yao, Z. Jiao, D. Ma, Extended-state-observer-based output feedback nonlinear robust control of hydraulic
systems with backstepping, IEEE Trans. Ind. Electron 61 (11) (2014) 6285–6293.
[3] J. Yao, Z. Jiao, D. Ma, A practical nonlinear adaptive control of hydraulic servomechanisms with periodic-like
disturbances, IEEE-ASME Trans. Mech 20 (6) (2015) 2752–2760.
[4] W.C. Sun, H.J. Gao, O. Kaynak, Finite frequency H-infinity control for vehicle active suspension systems, IEEE
Trans. Contr. Syst. Technol. 19 (2) (2011) 416–422.
[5] Z. Chen, B. Yao, Q.F. Wang, mu-Synthesis-based adaptive robust control of linear motor driven stages with
high-frequency dynamics: a case study, IEEE-ASME Trans. Mech 20 (3) (2015) 1482–1490.
[6] W.C. Sun, H.J. Gao, O. Kaynak, Vibration isolation for active suspensions with performance constraints and
actuator saturation, IEEE-ASME Trans. Mech 20 (2) (2015) 675–683.
[7] J. Yao, Z. Jiao, D. Ma, L. Yan, High-accuracy tracking control of hydraulic rotary actuators with modeling
uncertainties, IEEE-ASME Trans. Mech 19 (2) (2014) 633–641.
[8] L.T. Wang, G.F. Gong, H.Y. Yang, X. Yang, D.Q. Hou, The development of a high-speed segment erecting
system for shield tunneling machine, IEEE-ASME Trans. Mech 18 (6) (2013) 1713–1723.
[9] P.A. Sente, F.M. Labrique, P.J. Alexandre, Efficient control of a piezoelectric linear actuator embedded into a
servo-valve for aeronautic applications, IEEE Trans. Ind. Electron 59 (4) (2012) 1971–1979.
[10] J. Yao, Z. Jiao, S. Han, Friction compensation for low velocity control of hydraulic flight motion simulator: a
simple adaptive robust approach, Chin. J. Aeronaut 26 (3) (2013) 814–822.
[11] K. Seki, M. Iwasaki, M. Kawafuku, H. Hirai, K. Kishida, Practical controller design of hybrid experimental
system for seismic tests, IEEE Trans. Ind. Electron 56 (3) (2009) 628–634.
[12] J. Yao, W. Deng, W. Sun, Precision motion control for electro-hydraulic servo systems with noise alleviation:
A desired compensation adaptive approach, IEEE/ASME Trans. Mechatronics (2017), doi:10.1109/tmech.2017.
2688353.
[13] A. Mohanty, B. Yao, Indirect adaptive robust control of hydraulic manipulators with accurate parameter esti-
mates, IEEE Trans. Contr. Syst. Technol. 19 (3) (2011) 567–575.
[14] J. Yao, W. Deng, Active disturbance rejection adaptive control of hydraulic servo systems, IEEE Trans. Ind.
Electron. (2017), doi:10.1109/tie.2017.2694382.
[15] B. Yao, F.P. Bu, J. Reedy, G.T.C. Chiu, Adaptive robust motion control of single-rod hydraulic actuators: theory
and experiments, IEEE-ASME Trans. Mech 5 (1) (2000) 79–91.
[16] Z. Chen, Y.J. Pan, J. Gu, Integrated adaptive robust control for multilateral teleoperation systems under arbitrary
time delays, Int. J. Robust Nonlin 26 (12) (2016) 2708–2728.
[17] M.T.S. Aung, R. Kikuuwe, Acceleration feedback and friction compensation for improving positioning perfor-
mance in systems with friction, in: 2015 American Control Conference, 2015, pp. 4798–4803.
X. Li et al. / Journal of the Franklin Institute 354 (2017) 5328–5349 5349

[18] J. Yao, W. Deng, Z. Jiao, Adaptive control of hydraulic actuators with LuGre model-based friction compensation,
IEEE Trans. Ind. Electron 62 (10) (2015) 6469–6477.
[19] K.A.J. Verbert, R. Toth, R. Babuska, Adaptive friction compensation: a globally stable approach, IEEE-ASME
Trans. Mech 21 (1) (2016) 351–363.
[20] W.J. Chen, K. Kong, M. Tomizuka, Dual-stage adaptive friction compensation for precise load side position
tracking of indirect drive mechanisms, IEEE Trans. Contr. Syst. Technol. 23 (1) (2015) 164–175.
[21] F. Wilhelm, T. Tamura, R. Fuchs, P. Mullhaupt, Friction compensation control for power steering, IEEE Trans.
Contr. Syst. Technol. 24 (4) (2016) 1354–1367.
[22] K.M. Lee, C.H. Lee, S. Hwang, J. Choi, Y.B. Bang, Power-assisted wheelchair with gravity and friction com-
pensation, IEEE Trans. Ind. Electron 63 (4) (2016) 2203–2211.
[23] M. Sun, Z. Wang, Y. Wang, Z. Chen, On low-velocity compensation of brushless DC servo in the absence of
friction model, IEEE Trans. Ind. Electron 60 (9) (2013) 3897–3905.
[24] C.C. De Wit, H. Olsson, K.J. Astrom, P. Lischinsky, A new model for control of systems with friction, IEEE
Trans. Autom. Contr 40 (3) (1995) 419–425.
[25] L. Lu, B. Yao, Q.F. Wang, Z. Chen, Adaptive robust control of linear motors with dynamic friction compensation
using modified LuGre model, Automatica 45 (12) (2009) 2890–2896.
[26] A. Zaafouri, C.B. Regaya, H.B. Azza, A. Châari, DSP-based adaptive backstepping using the tracking errors
for high-performance sensorless speed control of induction motor drive, ISA Trans. 60 (2015) 333–347.
[27] R. Rashad, A. Aboudonia, A. El-Badawy, A novel disturbance observer-based backstepping controller with
command filtered compensation for a MIMO system, J. Franklin Inst. 353 (16) (2016) 4039–4061.
[28] C.B. Regaya, A. Zaafouri, A. Chaari, Speed sensorless indirect field-oriented of induction motor using two type
of adaptive observer, in: International Conference on Electrical Engineering and Software Applications, 2013,
pp. 1–5.
[29] F.S. Ahmed, S. Laghrouche, M. Harmouche, Adaptive backstepping output feedback control of DC motor
actuator with friction and load uncertainty compensation, Int. J. Robust Nonlin 25 (13) (2015) 1967–1992.
[30] L. Xu, B. Yao, Adaptive robust control of mechanical systems with nonlinear dynamic friction compensation,
Int. J. Control 81 (2) (2008) 167–176.
[31] A Ben-Israel, T.N.E. Greville, Generalized Inverses: Theory and Applications, Wiley, New York, 1974.
[32] B. Xian, M.S. de Queiroz, D.M. Dawson, M.L. McIntyre, A discontinuous output feedback controller and
velocity observer for nonlinear mechanical systems, Automatica 40 (4) (2004) 695–700.
[33] S.A. Ali, A. Christen, S. Begg, N. Langlois, Continuous–discrete time-observer design for state and disturbance
estimation of electro-hydraulic actuator systems, IEEE Trans. Ind. Electron 63 (7) (2016) 1.

Anda mungkin juga menyukai