Вы находитесь на странице: 1из 8

Observer-based cascade control of a 6-DOF parallel hydraulic manipulator

in joint space coordinate


Yangjun Pi
*
, Xuanyin Wang
The State Key Laboratory of Fluid Power Transmission and Control, Zhejiang University, Hangzhou 310027, China
a r t i c l e i n f o
Article history:
Received 4 November 2009
Accepted 3 July 2010
Keywords:
Disturbance observer
Parallel manipulator
Hydraulic actuation
Cascade control
a b s t r a c t
This paper addresses the joint space control problem of a 6-DOF (degree of freedom) parallel hydraulic
manipulator. High precision motion control of a six-degree parallel manipulator is hardly achieved due
to the existence of uncertain payload and other disturbance such as coupling force. A disturbance obser-
ver for this parallel manipulator is rst constructed to estimate and compensate the unknown distur-
bance. A cascade control algorithm is then applied to separate the hydraulic dynamics from the
mechanical part, which can mask the hydraulic dynamics with an inner loop. With such a control struc-
ture, known control design methods within the area of manipulator control can be directly used in the
outer loop. In this approach, the complex dynamics and direct kinematics of the parallel manipulator
are not required and acceleration feedback is also avoided. Experimental results are presented to show
the effectiveness of the proposed scheme.
2010 Elsevier Ltd. All rights reserved.
1. Introduction
The original 6-DOF parallel manipulator was rst developed in
1954 [1], and then became widely used in industry applications,
especially in motion simulator, due to its high power-to-weight ra-
tios, great stiffness and rapid responses [25]. However, the control
of parallel hydraulic manipulators is challenging as its dynamics is
highly nonlinear. First, the mechanical dynamics of the parallel
manipulators are much more complex than a single chain system,
which causes coupling force disturbance among the actuators. Sec-
ond, the payload or the external environment of the manipulator is
uncertain in some situations, for example, parallel manipulator is
subject to external disturbances which cannot be prior known.
Thirdly, hydraulic actuator dynamics are highly nonlinear due to
the phenomena such as nonlinear servo valve ow-pressure char-
acteristics, directional change of valve opening, valve overlap, and
position dependent dynamics [68]. Furthermore, hydraulic actua-
tors cannot be modeled as force/torque sources, it resemble veloc-
ity sources. Therefore, controllers that have been designed for
manipulator control, assuming the capability of setting actuator
force/torque, cannot be applied here [4]. The controllers that have
successfully incorporated actuator dynamic models can signi-
cantly improve the control performance [9]. However, it is shown
that the inclusion of hydraulic actuators increases the order of
the system [10].
Parallel manipulator control scheme mainly stems from two
frameworks [11,12]. The rst control scheme is called operational
space control scheme [13], and the second one is joint space con-
trol scheme. The former is based on the dynamics described by
the operational space. However, it needs the information from a
6-DOF sensor to obtain the pose and the displacement of the mov-
ing platform. Otherwise, it requires direct kinematics or other
methods to estimate the pose and displacement of the moving
platform. The joint space controller is designed to make the actual
actuator lengths conform to the desired lengths computed from
the command of the Cartesian coordinates of the moving platform
by inverse kinematics. This control scheme is much simple as it
does not need the information of the moving platform. However,
as introduced above, the coupling force disturbance and external
payload disturbance signicantly affect the performance of the
system. Additional force sensors can partly solve this problem,
but it does bring additional cost.
Disturbance observers which have been used in robotic manip-
ulator control for a long time can be used to estimate and compen-
sate the unknown disturbance [1416]. However, the analysis or
design of most existing disturbance observers, is based on linear
control theory, and the validity of using linear analysis and synthe-
sis techniques may be doubtful, since that the manipulators are
highly nonlinear and coupled system [17]. Recently some works
using nonlinear disturbance observer are reported [9,1822]. An
approach for designing nonlinear disturbance observer for serial
two-link robot manipulators was proposed by Chen et al. [17].
Nikoobin and Haghighi [23] extended it to general serial n-link
manipulators. In this paper, the nonlinear disturbance observer
0957-4158/$ - see front matter 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.mechatronics.2010.07.002
* Corresponding author. Tel.: +86 0571 87951271x6205; fax: +86 0571
87951941.
E-mail address: zjuppbird@gmail.com (Y. Pi).
Mechatronics 20 (2010) 648655
Contents lists available at ScienceDirect
Mechatronics
j our nal homepage: www. el sevi er. com/ l ocat e/ mechat r oni cs
which is designed for the parallel manipulators is simpler than the
one for the serial manipulators.
Hydraulic actuators have many distinct advantages such as the
ability to produce large forces at high speeds, high durability and
high power-to-weight ratios [24]. However, the dynamics of
hydraulic actuators is much more complex than that of electrical
actuators. To solve this problem, cascade control method has been
presented by Sepehri et al. [10] and Heintze et al. [2527]. Cascade
controller consists of two parts, namely the inner loop and the out-
er loop. It masks the hydraulic dynamics with the inner loop, and
then the outer loop does not concern the hydraulic dynamics. Koe-
kebakker [28] introduced cascade control to model based control
of a ight simulator motion system, the relatively complex dynam-
ics of Stewart platform was involved in the control design. Guo [2]
also presented a cascade position tracking controller to control a
hydraulic parallel robot manipulator. It presented a sliding mode
control method to reject external disturbance and assumed that
the estimation error of external disturbance is bounded by some
known function. However, the external disturbance cannot be
prior known in some situations.
This paper integrates cascade control and nonlinear disturbance
observer technique to design the controller for a 6-DOF hydraulic
parallel manipulator in joint space coordinate. By using a joint
space control scheme, the control problem of parallel manipulator
comes down to controlling the length of each actuator individually.
Cascade control that masks the hydraulic dynamics in an inner
loop, take into account not only the mechanical dynamics, but also
the hydraulic dynamics. Feedback linearization is adopted to com-
pensation the hydraulic nonlinear in the inner loop. It provides the
hydraulic actuator the characteristic of force generator. With such
a cascade control structure, some existing nonlinear control meth-
ods can be directly employed in the outer loop. A nonlinear distur-
bance observer is present to estimate the uncertain external
disturbance for the actuators of the parallel manipulator. The prop-
erties of individual parts of the proposed controller are already gi-
ven in literature, but an integral approach, having the specic
requirements of hydraulic parallel manipulator, is still lacking.
Experiments are carried out on a hydraulic 6-DOF parallel platform
and the performance improvements of the experimental results
demonstrate the effectiveness of the proposed approach.
2. Modeling
The 6-DOF hydraulic parallel manipulator consisted of a moving
platform, a base platform, six hydraulic actuators, and 12 universal
joints. The base platform and moving platform are connected by
the six hydraulic actuators, each of which has two universal joints.
A
i
, B
i
(i = 1, . . . , 6) are the coordinate of the universal joints of each
platform. The moving coordinate is O
p
-X
p
Y
p
Z
p
, and the base inertia
coordinate is O
g
-X
g
Y
g
Z
g
. The main structure parameters of the plat-
form are: the radius of the base platform R
0
, the radius of the mov-
ing platform ratio R
1
, the angle of two neighbor universal joints of
each platform h
0
, h
1
, as shown in Fig. 1.
2.1. Kinematics
The generalized coordinate vector, which describes the position
and orientation of the moving platform, can be dened as
X = [x, y, z, w
x
, w
y
, w
z
]
T
in the Cartesian space. The length vector of
six actuators is given by q = [q
1
, q
2
, . . . , q
6
]
T
. Inverse kinematics
concerns the determination of the length vector of six actuators
corresponding to a given generalized coordinate vector of moving
platform. Considering the ith actuator of the parallel manipulator,
as shown in Fig. 1, the length vector of ith actuator can be de-
scribed by
q
i
Rb
i
c a
i
1
where c = [x, y, z]
T
denotes the position of the point O
p
with respect
to the base inertia coordinate. Vector b
i
denotes the position of the
point B
i
with respect to the moving coordinate. Vector a
i
denotes
the position of the point A
i
with respect to the base inertia coordi-
nate, R represents the rotation matrix. The detailed expressions for
the rotation matrix can be found in [2,29].
The length of an actuator can be computed by the Euclidean
norm as
q
i
kq
i
k 2
2.2. Dynamics of hydraulic actuator
The six hydraulic actuators of the manipulator are controlled by
six hydraulic servo valves. Fig. 2 is the schematic of a single-rod
hydraulic servo system. It exhibit signicant nonlinear actuator
dynamics. In the following, the nonlinear dynamic model will be
gured out.
Taking leakage and compressibility into consideration, the
dynamics of cylinder oil ow can be written as follows [30]:
V
1
x
L

_
P
1
=b
e
A
1
_ x
L
Q
1
C
t
P
1
P
2
3
V
2
x
L

_
P
2
=b
e
A
2
_ x
L
Q
2
C
t
P
1
P
2
4
Fig. 1. Schematic diagram of a 6-DOF parallel manipulator.
Y. Pi, X. Wang / Mechatronics 20 (2010) 648655 649
where x
L
is the displacement of the piston, b
e
is the bulk modulus of
the uid, C
t
is the internal leakage coefcient, P
1
and P
2
are the pres-
sures inside the two chambers of the cylinder, Q
1
and Q
2
are the
uid ow rate of the two chambers of the cylinder. V
1
(x
L
) and
V
2
(x
L
) are the total uid volumes of the two sides of the cylinder,
and given as
V
1
x
L
V
10
A
1
x
L
5
V
2
x
L
V
20
A
2
L x
L
6
where L is the piston stroke, V
10
and V
20
are the initial uid volumes
of the two sides of the cylinder, A
1
and A
2
are the area of the two
sides of the piston.
As the volumes of the two chambers of the cylinder depend on
the position of piston, the dynamics of hydraulic cylinder is posi-
tion dependent. The stiffness of the hydraulic actuator is position
dependent [8].
The displacement of the piston is dened as
x
L
q q
0
7
where q
0
is the initial length of the actuators.
The uid ow rate of the two chambers of the cylinder are re-
lated to the spool valve displacement of the servo valve x
v
and gi-
ven as
Q
1
k
q
x
v

DP
1
p
; DP
1

P
s
P
1
x
v
0
P
1
P
r
x
v
< 0

8
Q
2
k
q
x
v

DP
2
p
; DP
2

P
2
P
r
x
v
0
P
s
P
2
x
v
< 0

9
where P
s
is the supply pressure, P
r
is the pressure of oil tank, which
is always zero, and k
q
is the ow gain coefcients of the servo valve
and given as
k
q
C
d
w

2=q
p
10
where w is the servo valve area gradient, C
d
is the discharge coef-
cient of cylinder and q is the oil density.
The forces balance equation for the asymmetric piston is given
by
A
1
P
1
A
2
P
2
mx
L
F
d
11
where m is the mass of the asymmetric piston. F
d
is the disturbance
force, which lumps payload force, friction force, gravity of piston
and other unmodeled disturbance. In Section 3, a nonlinear distur-
bance observer will be designed to estimate and compensate it.
For the completeness of the theory development, servo valve
dynamics and transmission line dynamics are introduced in the
following. In some literature, the servo valve opening x
v
is related
to the control input u by a rst-order system. In [8], a ve-order
dynamics was considered to describe the three-stage servo valve
dynamics. For long stroke hydraulic actuator, relatively long trans-
mission lines are present between the valve and the actuator
chambers, inducing badly damped resonances in the high fre-
quency range [8,27].
In this paper, the parallel manipulator work in the low fre-
quency range. For simplicity, the servo valve dynamics are ne-
glected and the servo valve opening x
v
is proportional to the
control input [2,9,31]. Then x
v
= k
a
u, where k
a
is the servo amplier
gain and u is the servo valve control input signal.
3. Controller design
3.1. Disturbance observer
In this section, a disturbance observer is designed to estimate
the disturbance force in (11). And (11) can be written as
F
d
A
1
P
1
A
2
P
2
mx
L
12
A disturbance observer is proposed as
_
b
F
d
k
o
b
F
d
k
o
F
d
k
o
b
F
d
k
o
A
1
P
1
A
2
P
2
mx
L
13
where
b
F
d
is the estimation of the disturbance force, and k
o
is a po-
sitive constant.
Dene the observer error as
e
d
F
d

b
F
d
14
Since there is no prior information about the derivative of the
disturbance F
d
, it is reasonable to assume that the disturbance
force in (11) varies slowly relative to the observer dynamics
[17,23], then
_
F
d
0 15
According to this assumption, the disturbance varies slowly rel-
ative to the observer dynamics [17]. It is reasonable to assume the
disturbance is constant during a short sampling period [23].
Combining (14) with the observer (13) yields
_ e
d

_
F
d

_
b
F
d
k
o
F
d

b
F
d
16
It can be expressed as below
_ e
d
k
o
e
d
0 17
The exponential convergence rate of the observer error can thus
be specied by choosing k
o
. However, the actuator acceleration
which is hardly obtained in this application appears in the distur-
bance observer. So, the disturbance observer given by (13) is not
practical to implement. To deal with this problem, an auxiliary var-
iable is dened as
k
b
F
d
p_ x
L
18
where p_ x
L
is the designed function.
Differentiating the auxiliary variable with respect to time yields
_
b
F
d

_
k
@p
@ _ x
L
x
L
19
Substituting for
_
b
F
d
from (13) into yields
_
k
@p
@ _ x
L
x
L
k
o
k p k
o
A
1
P
1
A
2
P
2
mx
L
20
By dening
@p
@ _ x
L
x
L
k
o
mx
L
21
then (20) can be written as
Fig. 2. Single-rod hydraulic servo system.
650 Y. Pi, X. Wang / Mechatronics 20 (2010) 648655
_
k k
o
k p k
o
A
1
P
1
A
2
P
2
22
The estimated disturbance can be obtained by
b
F
d
k p_ x
L
23
Comparing the two observers given by (13) and (23), the actu-
ator acceleration is not needed in the observer given by (23). How-
ever, it has two design parameters k
o
and p, which are not
independent and related to each other by (21). So, it must be
appropriately chosen to guarantee the asymptotical stability of this
observer. In this application, the function p_ x
L
is simpler than that
in [17,23].
Choosing
p c_ x
L
24
then
c k
o
m 25
The Lyapunov function candidate for the observer can be choos-
ing as
V me
2
d
26
Differentiating the Lyapunov function with respect to time
yields
_
V 2me
d
_ e
d
27
By combing (17) gives
_
V 2k
o
me
2
d
0 28
Then, the stability of the observer is guaranteed.
3.2. Inner-loop control
The aim of inner-loop control is to make the net force of the
hydraulic uid on the piston tracking the desired force. Heintze
et al. [8,2528] introduced an inner loop controller which includes
nonlinearity compensation. In this section, the inner loop control-
ler is designed based on the feedback linearization [27,32].
The net force of the hydraulic uid on the piston is given by:
F
p
A
1
P
1
A
2
P
2
29
Differentiating the uid force on the piston with respect to time
yields
_
F
p
A
1
_
P
1
A
2
_
P
2
30
After considering the hydraulic dynamics (3)(10), we have
_
F
p
f gu 31
where
f _ x
L
b
e
A
2
1
V
1

A
2
2
V
2
!

A
1
C
t
P
1
P
2

V
1

A
2
C
t
P
1
P
2

V
2

b
e
32
g
A
1
k
q

DP
1
p
V
1

A
2
k
q

DP
2
p
V
2

b
e
k
a
33
Choose control input u as
u
1
g
v f 34
where
v
_
F
pd
k
i
F
pd
F
p
35
and k
i
is a positive force error gain, F
pd
is the desired force input of
inner loop. Combining (31) and (34) yields the following equation:
_ e
F
k
i
e
F
0 36
where e
F
F
p
F
pd
.
It shows that the force error e
F
?0 with time constant 1/k
i
e
F
t e
k
i
t
e
F
0 37
This inner loop controller gives the hydraulic actuator the char-
acter of a force generator, where the net force of the hydraulic uid
on the piston F
p
tracks the desired force F
pd
with rst order dynam-
ics. The position dependent dynamics is compensated by this inner
loop controller.
In this paper, the servo valve dynamics and transmission line
dynamics are neglected. These dynamics can simply be added if
they appear to be essential [27].
Fig. 3. Controller implementation block diagram.
Y. Pi, X. Wang / Mechatronics 20 (2010) 648655 651
3.3. Outer loop control
In the above section, the inner loop controller is designed. How-
ever, the expression of the desired force F
pd
in (35) is still unknown.
The modeling error of hydraulic actuator which is used in inner-
loop control exists. Hence, the outer loop controller must be robust
to handle the modeling error and other uncertainties. Sliding mode
controllers are attractive for their capability to deal with uncer-
tainties. In the following, the desired net force will be derived on
the basis of the mechanical dynamics of hydraulic actuator using
the sliding mode control.
For simplicity, the forces balance equation for the asymmetric
piston (12) can be written as
x
L
w bF
p
38
where
w F
d
=m 39
b 1=m 40
Function w is not completely known but can be written as
w ^ w Dw, where ^ w is the nominal part and Dw is the uncertain
part, which is bounded by a known function f, i.e.,
jDwx; tj fx; t 41
^ w
b
F
d
=m 42
Obviously, the uncertain part Dw depends on the observer error
of the disturbance observer. As we know, prior knowledge of the
upper bounds of the disturbances is required to design a sliding
mode controller. However, the bounds of disturbances cannot be
prior known in some situations. By using disturbance observers,
only the bounds of observer errors of the disturbance observers
are needed to design the controller.
Fig. 4. The experimental setup.
652 Y. Pi, X. Wang / Mechatronics 20 (2010) 648655
The control objective is to make the displacement of the piston
x
L
track a desired trajectory x
d
asymptotically.
Let tracking error
~x x
L
x
d
43
Dene the sliding mode surface
sx; t
_
~x k~x 44
where k is a positive constant.
Differentiation of s with respect to t
_ s x
L
x
d
k
_
~x w bF
p
x
d
k
_
~x 45
Dene the Lyapunov function candidate as:
Vs
1
2
s
2
46
Differentiation of V
_
V s_ s sfw bF
p
x
d
k
_
~xg 47
The control law is designed as follows:
F
pd

1
b
^ u v 48
where
^ u w x
d
k
_
~x 49
v ksat
s
/

50
where / is the boundary layer of sliding mode, and sat(D) is the sat-
uration function, which can be formally dened as
satD
1; D > 1;
D; jDj 1;
1; D < 1:
8
>
<
>
:
51
If the discontinues item k is big enough, the sliding mode con-
dition will be match [32]. So, k is dened as
k f g 52
then
1
2
d
dt
s
2
gjsj 53
where g is a strictly positive constant.
The totally control scheme is shown in Fig. 3.
4. Experiment results
The proposed controller is experimentally evaluated on the
6-DOF parallel manipulator in the laboratory (see Fig. 4). The par-
allel manipulator is driven by six single-rod hydraulic cylinders
with a piston diameter of 63.5 mm and a rod diameter of
43.5 mm, and a full stroke of 500 mm. Each cylinder is controlled
by one servo valve (Moog J661-301) manufactured by Moog Indus-
trial Corporate.
The hardware conguration of the control system consists of
two PC-compatible computers: a host and a target. The host and
target PC are directly connected using a cross-over unshielded
twisted pair cable. All analog measurement signals are fed back
to a target PC through four plugged-in DAQ cards. The four DAQ
cards consist of two PCI1716Ls, a PCL812 and a PCL726 from the
Advantech Company. To attenuate the effects of noise, all mea-
sured signals are processed through a low-pass lter. The proposed
controller is implemented on the target PC, which is independent
of the host PC and must be booted using a special boot disc to load
the Matlab xPC Target Real-time Kernel. So the real-time applica-
tion is independent of the operating system and the control system
is more reliable and stable.
The load disturbance is directly measured by force sensor,
which is installed at the end of the piston, see Fig. 4. It can be used
Fig. 5. Position tracking of one of the actuators.
Fig. 6. Position tracking errors of one of the actuators (x
d
x
L
).
Fig. 7. Comparison result of the disturbance force.
Y. Pi, X. Wang / Mechatronics 20 (2010) 648655 653
to verify the effectiveness of the proposed nonlinear disturbance
observer. The disturbances which are given by the moving plat-
form can be measured by these force sensors. However, these dis-
turbances such as friction in the actuator cannot be obtained by
these force sensors. The control parameters used in the proposed
cascade controller with nonlinear disturbance are k
o
60; k
i

30; k 1; / 0:3 and g = 10.
The experimental results for the proposed cascade controller
with the nonlinear disturbance observer (Proposed), a P with speed
feed forward controller (Feedfoward) and a conventional PID
controller (PID) are compared. The reference trajectory is
z = 1540 + 50 sin (pt/8) in the operational space. The compared
experimental results in the joint space are shown in Figs. 5 and
6. In order to save publishing pages, only one joint result are pres-
ent here, while similar performance is observed for the other
joints. The results show that the performance is improved by using
the proposed controller. Figs. 7 and 8 give the comparison result of
one actuator between the estimation of disturbance of the observer
and the measure value of the sensor. Note that due to the friction
in the actuator and the gravity, the estimation of disturbance force
is not equal with the measured force by the force sensor. The dis-
turbance observer is more effective than force sensor in distur-
bance estimation. Figs. 9 and 10 show the net force tracking
behavior in the inner loop. In this case z = 1540 + 50 sin (1.2pt)
and the tracking result is given in Fig. 11. It shows that the tracking
Fig. 9. Inner loop tracking behavior.
Fig. 10. Inner loop tracking error.
Fig. 8. Force errors between disturbance observer and force sensor.
Fig. 11. Position tracking of one of the actuators.
Fig. 12. Position tracking errors (x
d
x
L
) of one of the actuators with the proposed
controller without disturbance observer.
654 Y. Pi, X. Wang / Mechatronics 20 (2010) 648655
error increased at a relatively higher frequency. So, the servo valve
dynamics may be not negligible at a relatively higher frequency.
The experimental result for the proposed cascade controller with-
out the nonlinear disturbance observer is given in Fig. 12. In Fig. 12
the prior estimate disturbances are given as 1200 N (Estimated 1),
1250 N (Estimated 2) and 1300 N (Estimated 3), respectively. It
shows that the prior knowledge of the disturbance have a signi-
cant impact on the control performance. Furthermore, if the bound
of disturbance is not prior known, the stability of this sliding mode
controller cannot guarantee.
5. Conclusions
This paper studies the joint spacecontrol problemof a 6-DOF par-
allel hydraulic manipulator. A cascade control algorithm is applied
to separate the hydraulic dynamics fromthe mechanical part, which
can mask the hydraulic dynamics with an inner loop. With such a
control structure, known control design methods within the area
of manipulator control can be directly used in the outer loop. A non-
linear disturbance observer is presented to estimate the unmodeled
disturbance. By using the disturbance observer, prior knowledge of
the upper bounds of the disturbances are not neededtodesigna slid-
ing mode controller. In this approach, the complex dynamics and di-
rect kinematics of the parallel manipulator are not required and
acceleration feedback is also avoided. Experiments are carried out
on a hydraulic 6-DOF parallel platform and the performance
improvements of the experimental results demonstrate the effec-
tiveness of the proposed approach. The proposed methodology can
be applied in almost all electro hydraulic systems with unknown
disturbances, such as construction machinery.
Acknowledgments
This work was supported by the National Natural Science
Foundation of China (Grant No. 50375139), and the New Century
excellent personage support plan of the Ministry of Education
(Grant No. NCET-04-0545). The authors greatly appreciate the
comments and suggestions by the reviewers.
References
[1] Stewart D. A platform with six degrees of freedom. Proc Inst Mech Eng, Part A: J
Power Energy 1965;180:37185.
[2] Guo HB, Liu YG, Liu GR, Li HR. Cascade control of a hydraulically driven 6-DOF
parallel robot manipulator based on a sliding mode. Control Eng Pract
2008;16:105568.
[3] Dasgupta B, Mruthyunjaya TS. The Stewart platform manipulator: a review.
Mech Mach Theory 2000;35:1540.
[4] Davliakos I, Papadopoulos E. Model-based control of a 6-dof electrohydraulic
StewartGough platform. Mech Mach Theory 2008;43:1385400.
[5] Merlet J. Parallel robots. Kluwer Academic Publishers; 2006.
[6] Guan C, Pan SX. Adaptive sliding mode control of electro-hydraulic system
with nonlinear unknown parameters. Control Eng Pract 2008;16:127584.
[7] Wijnheijmer F, Naus G, Post W, Steinbuch M, Teerhuis P. Modelling and LPV
control of an electro-hydraulic servo system. In: IEEE International conference
on control applications, Munich, Germany; 2006. p. 311621.
[8] van Schothorst G. Modelling of long-stroke hydraulic servo-systems for ight
simulator motion control and system design. PhD-thesis, Delft University of
Technology; 1997.
[9] Sirouspour MR, Salcudean SE. Nonlinear control of hydraulic robots. IEEE Trans
Rob Autom 2001;17:17382.
[10] Sepehri N, Dumont G, Lawrence P, Sassani F. Cascade control of hydraulically
actuated manipulators. Robotica 1990;8:20716.
[11] Kim DH, Kang JY, Lee KI. Robust tracking control design for a 6 DOF parallel
manipulator. J Rob Syst 2000;17:52747.
[12] Lee SH, Song JB, Choi WC, Hong DH. Position control of a Stewart platform
using inverse dynamics control with approximate dynamics. Mechatronics
2003;13:60519.
[13] Kim HS, Cho YM, Lee KI. Robust nonlinear task space control for 6 DOF parallel
manipulator. Automatica 2005;41:1591600.
[14] Jezernik K, Curk B, Harnik J. Observer-based sliding mode control of a robotic
manipulator. Robotica 2009;12:4438.
[15] Chan SP. A disturbance observer for robot manipulators with application to
electronic components assembly. IEEE Trans Ind Electron 1995;42:48793.
[16] Kim BK, Chung WK, Youm Y. Robust learning control for robot manipulators
based on disturbance observer. In: 22nd International conference on industrial
electronics, control, and instrumentation, Taipei; 1996. p. 127682.
[17] Chen W, Ballance D, Gawthrop P, OReilly J. A nonlinear disturbance observer
for robotic manipulators. IEEE Trans Ind Electron 2000;47:9328.
[18] Kishore WCA, Sen S, Ray G. Disturbance observer based performance
improvement of H-innity controller. In: IEEE region 10 colloquium/3rd
international conference on industrial and information systems, Kharagpur,
India; 2008. p. 28590.
[19] Zhu L, Lu P, Jing ZL, Hu SQ. A new robust adaptive control scheme for non-
afne nonlinear systems based on SHLNN disturbance observer. In: 4th
International symposium on management engineering, Kitakyushu, Japan;
2007. p. 163342.
[20] Son YI, Shim H, Jo NH, Kim SJ. Design of disturbance observer for non-
minimum phase systems using PID controllers. In: Annual conference on the
society of instrument and control engineers, Takamatsu, Japan; 2007. p. 196
201.
[21] Kravaris C, Sotiropoulos V, Georgiou C, Kazantzis N, Xiao M, Krener A.
Nonlinear observer design for state and disturbance estimation. Syst Control
Lett 2007;56:7305.
[22] Chen X, Komada S, Fukuda T. Design of a nonlinear disturbance observer. IEEE
Trans Ind Electron 2000;47:42937.
[23] Nikoobin A, Haghighi R. Lyapunov-based nonlinear disturbance observer for
serial n-link robot manipulators. J Intell Rob Syst 2009;55:13553.
[24] Yao B, Bu FP, Reedy J, Chiu GTC. Adaptive robust motion control of single-rod
hydraulic actuators: theory and experiments. IEEE/ASME Trans Mechatron
2000;5:7991.
[25] Heintze J, Van der Weiden A. Inner-loop design and analysis for hydraulic
actuators, with an application to impedance control. Control Eng Pract
1995;3:132330.
[26] Heintze J, Peters RM, Weiden AJJVD. Cascade Dp and sliding mode for
hydraulic actuators. In: Proceedings of the 3rd European control conference,
Rome, Italy; 1995. p. 14717.
[27] Heintze J. Design and control of a hydraulically actuated industrial brick laying
robot. PhD-thesis, Delft University of Technology; 1997.
[28] Koekebakker S. Model based control of a ight simulator motion system. PhD-
thesis, Delft University of Technology; 2001.
[29] Ghobakhloo A, Eghtesad M, Azadi M. Position control of a StewartGough
platform using inverse dynamics method with full dynamics. In: 9th IEEE
international workshop on advanced motion control, Istanbul, Turkey; 2006. p.
505.
[30] Merritt H. Hydraulic control systems. Wiley; 1967.
[31] Sohl GA, Bobrow JE. Experiments and simulations on the nonlinear control of a
hydraulic servosystem. IEEE Trans Control Syst Technol 1999;7:23847.
[32] Slotine J, Li W. Applied nonlinear control. Prentice Hall; 1991.
Y. Pi, X. Wang / Mechatronics 20 (2010) 648655 655

Вам также может понравиться