You are on page 1of 9

Trajectory tracking control of a 6-DOF hydraulic parallel robot manipulator

with uncertain load disturbances


Yangjun Pi
n
, 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 2 June 2009
Accepted 18 November 2010
Available online 8 December 2010
Keywords:
6-DOF
Hydraulic actuators
Parallel robot manipulator
Force sensors
Uncertain load disturbances
a b s t r a c t
This study addresses the trajectory tracking control of a 6-DOF (degrees of freedom) hydraulic parallel
robot manipulator with uncertain load disturbances. As load disturbances are the main external
disturbances of the parallel robot manipulators and have a signicant impact on system tracking
performance, many researchers have been devoted to synthesize advanced control methods for
improving the system robustness under the assumption that load disturbances are bounded. However,
load disturbances are uncertain and vary in a large range in real situation happening in most hydraulic
parallel robot manipulators, which is opposed to the assumption. In this paper, the load disturbances are
directly measured by force sensors. Then a sliding mode control with discontinuous projection-based
adaptation laws is proposed to improve the tracking performance of the parallel robot manipulator.
Simulations and experiments with typical desired trajectory are presented, and the results show that
good tracking performance is achieved in the presence of uncertain load disturbances.
& 2010 Elsevier Ltd. All rights reserved.
1. Introduction
Parallel robot manipulators are widely used in the entertainment
andmachinetool industries, especiallyinmotionsimulators (Davliakos
& Papadopoulos, 2008; Sirouspour & Salcudean, 2001). They have
many distinct advantages compared with the serial robot manipula-
tors, suchas the fast response speed, highsystemstiffness, anda higher
force-to-weight ratio (Dasgupta & Mruthyunjaya, 2000; Davliakos &
Papadopoulos, 2008; Kim, Kang, & Lee, 2000; Sirouspour & Salcudean,
2001). Many researchers have been studying on those robot manip-
ulators since the original GoughStewart parallel mechanism was
developed by Gough and Stewart in 1954 and 1965, respectively
(Dasgupta & Mruthyunjaya, 2000; Guo, Liu, Liu, & Li, 2008; Lee, Song,
Choi, & Hong, 2003; Stewart, 1965; Ting, Chen, & Jar, 2004). A number
of studies onthe kinematics anddynamics of those robot manipulators
have been published (Baron & Angeles, 2000; Bonev & Ryu, 2000; Gao,
Lei, Liao, & Zhang, 2005; Guo & Li, 2006; Lebret, Liu, & Lewis, 1993;
Merlet, 2004; Taghirad & Nahon, 2008). Accurate trajectory tracking
control for parallel robot manipulators is a key systemrequirement, as
these devices must often follow prescribed motion (Ghobakhloo,
Eghtesad, &Azadi, 2006; Tafazoli, de Silva, &Lawrence, 1998). Tracking
control of parallel robot manipulators has been approached using both
linear and nonlinear control laws (Sirouspour & Salcudean, 2001; Sohl
&Bobrow, 1999). Highperformance control strategies cansignicantly
improve the tracking performance (Dasgupta & Mruthyunjaya, 2000;
Sirouspour & Salcudean, 2001).
In general, parallel robot manipulators can be driven by electrical
or hydraulic actuators. The actuator dynamics may be neglected
in many cases due to the fast time constants of electrically
driven actuators. Unlike the electrical actuators that resemble force
sources, hydraulic actuators resemble velocity sources (Davliakos &
Papadopoulos, 2008; Sirouspour & Salcudean, 2001), and cannot
accurately apply forces or torques over a signicant dynamic range
(Davliakos & Papadopoulos, 2008). Considering the hydraulic actua-
tor dynamics, controllers that have been designed for electrical
driven robot manipulator control cannot be directly used here.
Hydraulic systems also have a number of characteristics which
complicate the development of high-performance controllers. The
dynamics of hydraulic systems are highly nonlinear due to the
phenomena suchas uidcompressibility, nonlinear servovalve ow-
pressure characteristics, and dead band due to the internal leakage
and hysteresis (Chen, Renn, &Su, 2005; Yao, Bu, Reedy, &Chiu, 2000;
Zhu & Piedboeuf, 2005). However, compared with electrical actua-
tors, hydraulic actuators have many distinct advantages such as the
ability to produce large forces at high speeds, the high durability and
stiffness, and rapid response (Chen et al., 2005; Yao et al., 2000).
Therefore, this paper will address the tracking control of hydraulic
robot manipulators.
There are two frameworks for the parallel robot manipulator
control scheme (Davliakos & Papadopoulos, 2008; Ghobakhloo
et al., 2006; Guo et al., 2008). The rst control scheme is called
operational space control scheme. The control scheme is based on
the dynamics described by the operational space and needs the
Contents lists available at ScienceDirect
journal homepage: www.elsevier.com/locate/conengprac
Control Engineering Practice
0967-0661/$ - see front matter & 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.conengprac.2010.11.006
n
Corresponding author. Tel.: +86 0571 87951271/6205;
fax: +86 0571 87951941.
E-mail address: zjuppbird@gmail.com (Y. Pi).
Control Engineering Practice 19 (2011) 185193
Cartesian coordinates of the moving platform. This information can be
obtained by 6-DOF Cartesian position and orientation sensors or by
other methods that can estimate the Cartesian coordinates of the
moving platform such as forward kinematics (Bonev & Ryu, 2000;
Merlet, 2004). However, the 6-DOF sensors may be very expensive or
impossible to nd at the desired accuracies. Forward kinematics of six-
degree parallel robot manipulator can be solved in a few milliseconds
due to the development of digital technology development, though it
relays on numerical methods. So, it is possible to use forward
kinematics in real time control. The second control scheme is called
joint space control scheme. This controller is designed to make the
actual actuator lengths conformto the desired lengths computed from
the command of the Cartesian coordinates of the moving platform by
inverse kinematics. There is certain relationship between Cartesian
coordinates of the moving platform and the six actuator lengths, as
these robot manipulators are rigidlinked. If the six actuator lengths are
appropriately controlled, the moving platform of the parallel robot
manipulator cantrack desiredtrajectories. Bothframeworks have their
advantages and disadvantages. The former appears to be preferable,
since it candirectly obtainthe tracking errors of the robot manipulator.
But the Cartesian coordinates of the moving platform is difcult to
obtain.
Load disturbances are the main external disturbances of the
parallel robot manipulators, and always impact on system perfor-
mance greatly. It is difcult to simultaneously guarantee the accuracy
androbustness for a controller, whenthe systemloadis uncertainand
varying. Alot of research studies design robust controller to deal with
the external disturbances, however, it depends on the assumptions
that the external disturbances are bounded. An adaptive control of
manipulators using uncalibrated joint-torque sensing was designed
for the control of electrically actuated serial manipulators (Farhad
Aghili & Mehrzad Namvar, 2005). It is easy to install extra force
sensors at the end terminal of actuators for parallel robot manip-
ulators. Then, the forces that are applied by the load and the moving
platform can be detected by extra force sensors. Load disturbances
are attenuated by this method, and the system is more robust.
Obviously, the joint space control scheme is preferable here. This
system can be considered as six single-rod hydraulic actuators. The
control of single-rod hydraulic actuators has been studied by many
researchers (Chen et al., 2005; Yao et al., 2000; Zhu & Piedboeuf,
2005). The two most common approaches developed to compensate
for the nonlinear behavior of hydraulic parallel robot manipulators
are adaptive control and sliding mode control (Bonchis, Corke, Rye, &
Ha, 2001; Guan & Pan, 2008; Kim et al., 2000; Slotine & Li, 1991).
Adaptive control is a valid method to overcome system parameter
variations, and sliding mode control has been an effective way for
controlling nonlinear systems with uncertainties, especially unmo-
deled uncertainties.
In this paper, the controller that combines the sliding mode
control method and some unknown parameters update laws is
developed. The sliding mode control contains discontinuous func-
tion, which can cause chattering phenomena. Chattering is usually
undesirable because it results in a high heat loss. In addition,
chattering may excite unmodeled high-frequency dynamics so as
to degrade the performance of the system. In order to avoid the
discontinuous function, a continuous approximation of the sliding
mode control is employed in the proposed controller.
To test the proposed control scheme, the proposed nonlinear
control with force sensors (NCFS), nonlinear control with force
observer (NCFO) and PID control strategies are compared by
simulation and experimentation. The results show that the pro-
posed nonlinear control strategy has a better performance when
compared with the other two control schemes.
The paper is organized as follows. System dynamics, including
parallel robot manipulator and hydraulic actuator dynamics are given
in Section 2. In Section 3, the controller is proposed. Simulation results
aregiveninSection4. Theexperimental results arepresentedinSection
5. Finally, conclusions are given in Section 6.
2. Robot modeling
2.1. Rigid body dynamics of six-degree parallel robot manipulator
The six-degree parallel robot manipulator consists of a base
platform, a moving platform, and six hydraulic actuators. The
moving platformis supported by six hydraulic actuators, see Fig. 1.
The rigid body dynamics of six-degree parallel robot manipulator
can be described as a second-order nonlinear differential equation,
which is given by
M

X V
_
X Gs
r
J
T
f
F 1
where XAR
6
is a vector of the platformgeneralizedcoordinates and
F AR
6
is the vector of the generalized applied force of the actuators.
MAR
66
is the robot manipulator mass matrix, VAR
66
contains
coriolis and centripetal terms and GAR
6
represents gravitational
effects. s
r
AR
6
is the load force effects. J
f
AR
66
is the force Jacobian
matrix of the system. q is a vector of the length of joints.
2.2. Dynamics of the hydraulic actuators
The schematic of a single-rod hydraulic actuator is shown in
Fig. 2. Each hydraulic actuator consists of an asymmetric piston, a
servo valve and sensors. The piston motion is obtained by
modulating the oil ow into and out of the cylinder chambers,
which are connected to a servo valve through cylinder ports. It
exhibits signicant nonlinear actuator dynamics. In the following,
the nonlinear dynamic model will be given.
The force balance equationfor the asymmetric pistonis givenby
A
1
P
1
A
2
P
2
m x
L
b_ x
L
F
d
2
where x
L
is the displacement of the piston, m is the mass of the
asymmetric piston, A
1
and A
2
are the area of the two sides of
the piston, P
1
and P
2
are the pressures inside the two chambers of the
cylinder, b is the coefcient of viscous friction, F
d
is the load force.
In this paper, the estimation of load force F
d
is obtained by two
methods. One is by force sensors; the other is by force observer.
Consider the systemdynamics described by (1); the force observer
dynamics are given as
^
F
d
J
T
f

1
M

X V
_
X G 3
Note that load force effect s
r
is not considered in the force
observer, so the observed value of the load force is not accurate in
the presence of uncertain load disturbances.
Fig. 1. Schematic diagram of a 6-DOF parallel robot manipulator.
Y. Pi, X. Wang / Control Engineering Practice 19 (2011) 185193 186
The estimation error of load force is given as
~
F
d
F
d

^
F
d
4
Taking leakage and compressibility into consideration, the
dynamics of cylinder oil ow can be written as follows (Merritt,
1967):
V
1
x
L

_
P
1
=b
e
A
1
_ x
L
Q
1
C
t
P
1
P
2
5
V
2
x
L

_
P
2
=b
e
A
2
_ x
L
Q
2
C
t
P
1
P
2
6
where b
e
is the bulk modulus of the uid, C
t
is the internal leakage
coefcient, 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
7
V
2
x
L
V
20
A
2
Lx
L
8
where L is the pistonstroke, V
10
andV
20
are the initial uid volumes
of the two sides of the cylinder. Q
1
and Q
2
are the uid ow rate of
the two chambers of the cylinder. They are related to the spool
valve displacement of the servo valve x
v
and given as
Q
1
k
q
x
v

DP
1
_
, DP
1

P
s
P
1
x
v
Z0
P
1
P
r
x
v
o0
_
9
Q
2
k
q
x
v

DP
2
_
, DP
2

P
2
P
r
x
v
Z0
P
s
P
2
x
v
o0
_
10
where P
s
is the supply pressure, P
r
is the pressure of oil tank,
whichis always zero, and k
q
is the owgaincoefcients of the servo
valve given as
k
q
C
d
w

2=r
_
11
where w is the servo valve area gradient, C
d
is the discharge
coefcient of the cylinder and ris the oil density.
Remark 1. For simplifying the controller design, the servo valve
model given by (9) and (10) is relatively crude but sufcient in this
paper. In some situation (i.e., the leakage ow between the valve
spool and body cannot be ignored), more evolved models should be
adopted (Eryilmaz &Wilson, 2001; Ferreira, De Almeida, &Quintas,
2002).
Remark 2. Reverse owmust be taken into account as the parallel
robot manipulator is inertially loaded. It should be avoided by
carefully designing in advance.
In some literatures, the dynamics of servo valve have been
incorporated in the controller design. Here, the valve dynamics are
neglected and the servo valve opening x
v
is proportional to the
control input, since a high-response servo valve is used. Then
x
v
k
a
u, where k
a
is the servo amplier gainand u is the servo valve
control input signal.
Dening the state variables as
x x
1
,x
2
,x
3
,x
4

T
9x
L
, _ x
L
,P
1
,P
2

T
12
The entire dynamics can be expressed in a state-space form as
_ x
1
x
2
_ x
2

1
m
A
1
x
3
A
2
x
4
bx
2
F
d

_ x
3

b
e
V
1
A
1
x
2
k
q
k
a

DP
1
_
uC
t
x
3
x
4

_ _
_ x
4

b
e
V
1
A
2
x
2
k
q
k
a

DP
2
_
uC
t
x
3
x
4

_ _
13
3. Control of the six-degree parallel robot manipulator
In general, the system is subjected to parametric uncertainties
due to the variations of b
e
, m, b, C
t
, k
q
, k
a
. The coefcient of viscous
friction is neglected in the following [i.e., let b0] to test the
robustness of the proposed control scheme. In order to use
parameter adaptation to reduce parametric uncertainties for an
improved performance, it is necessary to linearly parameterize the
state-space equation (13) in terms of a set of unknown parameters.
Let us dene the unknown parameter set h y
1
,y
2
,y
3
,y
4
_
as
y
1
b
e
=m, y
2
b
e
C
t
=m, y
3
1=m, y
4
b
e
k
q
k
a
=m. The state-space
Eq. (13) can be linearly parameterized in terms of h as
x

L
y
1
A
2
1
V
1

A
2
2
V
2
_ _
x
2
y
2
A
1
V
1

A
2
V
2
_ _
x
3
x
4
y
3
_
F
d
y
4
A
1
V
1

DP
1
_

A
2
V
2

DP
2
_
_ _
u 14
Dening the estimation value of h as
^
h
^
y
1
,
^
y
2
,
^
y
3
,
^
y
4
, and the
estimation error given as
~
y
i
y
i

^
y
i
, i 1,2,. . .,4 15
In the following, the sliding mode control is presented, and the
adaptation laws are used to compensate for the uncertain
parameters.
For simplicity, Eq. (14) can be written as
x

L
f x,tgxu 16
where
f y
1
A
2
1
V
1

A
2
2
V
2
_ _
x
2
y
2
A
1
V
1

A
2
V
2
_ _
x
3
x
4
y
3
_
F
d
17
g y
4
A
1
V
1

DP
1
_

A
2
V
2

DP
2
_
_ _
18
Function f is not completely known but can be written
asf
^
f Df , where
^
f is the nominal part and Df is the uncertain
part, which is bounded by a known functionF, i.e.,
9Df x,t9rFx,t 19
^
f
^
y
1
A
2
1
V
1

A
2
2
V
2
_ _
x
2

^
y
2
A
1
V
1

A
2
V
2
_ _
x
3
x
4

^
y
3
_
^
F
d
20
The control gain g is unknown but conned to a certain constant
range. i.e.,
0ob
min
rgx rb
max
21
The estimated value of gx is given as
^ g
^
y
4
A
1
V
1

DP
1
_

A
2
V
2

DP
2
_
_ _
22
The control objective is to make the system state x
L
track a
desired trajectory x
Ld
asymptotically.
Fig. 2. Single-rod hydraulic servo system.
Y. Pi, X. Wang / Control Engineering Practice 19 (2011) 185193 187
Considering the tracking error as
~ x
L
x
L
x
Ld
23
Dening the sliding mode surface
sx; t
d
dt
l
_ _
2
~ x
L


~ x
L
2l
_
~ x
L
l
2
~ x
L
24
where l is a positive constant.
Differentiation of s with respect to t
_ s x

L
x

Ld
2l

~ x
L
l
2
_
~ x
L
f gux

Ld
2l

~ x
L
l
2
_
~ x
L
25
Dening the Lyapunov function candidate as
Vs
1
2
s
2
26
Differentiation of V
_
V s_ s s y
1
A
2
1
V
1

A
2
2
V
2
_ _
x
2
y
2
A
1
V
1

A
2
V
2
_ _
x
3
x
4
y
3
_
F
d
_
y
4
A
1
V
1

DP
1
_

A
2
V
2

DP
2
_
_ _
ux

Ld
2l

~ x
L
l
2
_
~ x
L
_
27
The control input u is designed as follows:
u
1
^ g
^ uv 28
where
^ u
^
f x

Ld
2l

~ x
L
l
2
_
~ x
L
29
v ksat
s
f
_ _
30
where k is the control gain, fis the boundary layer of sliding mode,
and satDis the saturation function, which can be formally dened
as
satD
1, D41,
D, 9D9r1,
1, Do1
_

_
31
In order to obtain the adaptation laws of parameters, the
following Lyapunov function is dened as
V
1
s
1
2
s
2

1
2

4
i 1
U
i
~
y
i
2
32
whereU
i
, i 1,2,. . .,4 are positive denite constants.
Taking into account Eqs. (15) and (27)(31), the time derivative
of V
1
is
_
V
1
s_ s

4
i 1
U
i
~
y
i
_
^
y
i

~
y
1
s
A
2
1
V
1

A
2
2
V
2
_ _
x
2
U
1
_
^
y
1
_ _

~
y
2
s
A
1
V
1

A
2
V
2
_ _
x
3
x
4
U
2
_
^
y
2
_ _

~
y
3
s
_
^
F
d
U
3
_
^
y
3

~
y
4
s
A
1
V
1

DP
1
_

A
2
V
2

DP
2
_
_ _
uU
4
_
^
y
4
_ _
ksat
s
f
_ _
sy
3
_
~
F
d
s
33
to make sure that
_
V
1
r0, the adaptation laws are chosen as
_
^
y
1
Proj
^
y
1
U
1
1
A
2
1
V
1

A
2
2
V
2
_ _
x
2
s
_ _
34
_
^
y
2
Proj
^
y
2
U
1
2
A
1
V
1

A
2
V
2
_ _
x
3
x
4
s
_ _
35
_
^
y
3
Proj
^
y
3
U
1
3
_
^
F
d
s 36
_
^
y
4
Proj
^
y
4
U
1
4
A
1
V
1

DP
1
_

A
2
V
2

DP
2
_
_ _
us
_ _
37
FunctionProj
^
y
i
Y
i
is a simple discontinuous projection, dened
as
Proj
^
y
i
Y
i

0 if
^
y
i
y
imax
and Y
i
40,
0 if
^
y
i
y
imin
and Y
i
o0,
Y
i
otherwise
_

_
38
It can be shown that for any adaptation functionD
i
, the projec-
tion mapping used in (34)(37) guarantees (Guan &Pan, 2008; Yao
et al., 2000)
1
^
y
i
Af
^
y
i
: y
imin
r
^
y
i
ry
imax
g
2
~
y
i
U
i
Proj
^
y
i
U
1
i
D
i
D
i
r0
39
Substituting the above adaptation laws (34)(37) into (33)
_
V
1
ksat
s
f
_ _
sy
3
_
~
F
d
s 40
Assumption 1. The measurement error of the force sensor is small
enough, i.e.,
~
F
d
0.
According to Assumption 1
_
V
1
ksat
s
f
_ _
sr0 41
The controller is proposedtoguarantee asymptotic convergence
to zero of the tracking error, i.e. ~ xt-0 as time t-1.
Remark 3. It has been shown that the time derivative of the
estimation error of load force plays an important role in the
Lyapunov function (40). If the load force varies relatively slowly,
a rough estimation of load force is sufcient for controller design.
Otherwise, the method proposed in this paper is an alternative.
4. Simulation results
The proposed control laws were simulated using Matlabs
Simulink package. The controller parameters of the proposed
nonlinear controller are designed asl 20,f0:1 andk 1:2.
The sample time is set as 0.002 s. The implementation block
diagrams of the controllers are shown in Fig. 3(a) and (b). Note
that q q
0
x
L
, where q
0
is the initial length of the joints.
In order to obtain the guidelines for experimentation, the
realistic model of the hydraulic parallel robot manipulator has
been used. The system parameters are selected based upon their
actual values and are given in Table 1.
The reference trajectory is chosen to be
x y 0:1sin
p
8
t
_ _
, z 1:6
c
x
c
y
c
z
5sin
p
8
t
_ _
_
42
Positions and angles are expressed in meters and degrees,
respectively.
Nonlinear control with force sensors (NCFS), nonlinear control
withforce observer (NCFO) andPIDcontrol strategies are compared
in the reference trajectory given by Eq. (42). In order to obtain the
performance and robustness with load disturbances, the external
disturbance given by Eq. (43) is introduced in each actuator.
F
d

0 t o60s
2500sinpt t Z60s
_
43
The load disturbances are expressed in Newton.
The tracking errors of a well-tuned PID controller are shown in
Figs. 4 and 5. The tracking errors of nonlinear controller with force
observer are given by Figs. 6 and 7. The position and orientation errors
Y. Pi, X. Wang / Control Engineering Practice 19 (2011) 185193 188
of nonlinear controller with force sensors are depicted in Figs. 8 and 9.
The simulation results show that the nonlinear controller with force
sensors yields the smallest tracking errors in all situations. It shows a
goodresponse of the nonlinear controller withforce sensors, especially
when the external disturbances are introduced. The nonlinear control
with force observer shows the same performance with the nonlinear
control with force sensors, when the load disturbances are bounded.
Whentheexternal disturbances are introduced, the performanceof the
nonlinear control withforceobserver is signicantlydegraded. And, the
well-tuned PID controller shows the worst performance among the
three control strategies.
To test the performance of the proposed adaptation control
laws, the original values of system parameters are set as h
3 10
7
,0,0:05,3. The real system parameters are designed as
h 3:45 10
7
,0,0:05,2:45. The bounds of system parameters are
designed as h
min
2 10
7
,0,0:03,1:5, and h
max
5 10
7
,0:2,
0:07,3:5. The adaptation rates are designed as ! 1 10
7
,1,
1000,1000. The parameter adaptation laws were activated after
t40 s. The estimates of y
1
and y
4
reach their boundaries during
some periods of the simulation as seen in Figs. 10 and 11.
5. Experimental results
The proposed control strategies were experimentally evaluated
using the 6-DOF parallel robot manipulator (see Fig. 12). The robot
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
Fig. 3. Controller implementation block diagrams. (a) controller with force sensors and (b) controller with force observer.
Table 1
The parameters of the parallel robot.
Parameter Description Value Unit
A
1
Area of the cylinder-end of the piston
4:03 10
3
m
2
A
2
Area of the rod-end of the piston
2:03 10
3
m
2
L Piston stroke 0.5 m
w Servo valve area gradient 0.008 m
b
e
Bulk modulus of the uid 690 MPa
P
s
Supply pressure 2 MPa
M
p
Mass of moving platform 550.3 kg
I
x
Inertia around x-axis 78.5 kg m
2
I
y
Inertia around y-axis 108.25 kg m
2
I
z
Inertia around z-axis 185.26 kg m
2
m Mass of piston 20 kg
Fig. 4. The position tracking errors with PID controller.
Y. Pi, X. Wang / Control Engineering Practice 19 (2011) 185193 189
full stroke of 500 mm. Each cylinder is controlled by a servo valve
(Moog J661-301) manufactured by Moog Industrial Corporate. The
integrated position transducer measures the position of the piston.
The dynamics of the valves were ignored. The actuator velocities
are estimated from the measured actuator positions. Pressure
sensors are installed on each chamber of the cylinder. The radii
of the base and moving platforms are 1 and 0.8 m, respectively. The
angle between the centers of the two pair of joints is 1201. The force
sensors are installed in the cylinders and platform, as connecter
(see Fig. 12). For simplicity, the initial values of the parameter
estimates were identied by off-line experiments.
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 of cables. All analog measured signals (the cylinder
position, the chamber pressures, the supplied pressure, and
the load forces) 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 inuence of noise, all measured signals are
processed through a low-pass lter. The proposed three controllers
are 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 application is inde-
pendent of the operating system and the control system is more
reliable and stable. By using this setup, a control frequency of
500 Hz is achieved.
To verify the effectiveness of the proposed control scheme, there
should be a comparative study between nonlinear controller
with force sensors and nonlinear controller with force observer.
Fig. 5. The orientation tracking errors with PID controller.
Fig. 6. The position tracking errors with NCFO.
Fig. 7. The orientation tracking errors with NCFO.
Fig. 8. The position tracking errors with NCFS.
Fig. 9. The orientation tracking errors with NCFS.
Fig. 10. The estimation errors of y
1
.
Y. Pi, X. Wang / Control Engineering Practice 19 (2011) 185193 190
However, the two control schemes show similar performance in
experiments, since the load is relatively small. So, only the com-
parative results of the experiments with nonlinear controller with
force sensors and well-tuned PID control are presented here. The
controller parameters are designed as l 20, f0:1 and k 3.
The tracking errors of well-tuned PID control and nonlinear
controller with force sensors in tracking the reference trajectory
x 0:1cospt=8; y 0:1sinpt=8 m are shown in Figs. 13 and 14,
respectively. The maximum tracking errors are 5 and 1.4 mm for
the PID controller and nonlinear controller with force sensors,
respectively. Similar results were obtained in the other coordi-
nates. For example, the tracking errors around x coordinate c
x
where c
x
2sinpt=8 degrees for well-tuned PID controller and
the nonlinear controller with force sensors are shown in Figs. 15
and 16, respectively. The force disturbances of each joint actuator
are given by Fig. 17. Note that due to the friction in the actuators
and the measured noise, the estimated parameters did not con-
verge to x values.
Despite the experimental conditions are different, the proposed
control scheme is also compared with a recent published study for
this kind of robot manipulators in literature (Guo et al., 2008). The
tracking error around x coordinate, where c
x
0:5sin2ptdegrees
for the proposed controller is shown in Fig. 18. The tracking
Fig. 11. The estimation errors of y
4
.
Windows
XP
(Simulink)
Host PC
E
t
h
e
r
n
e
t
Target PC
Matlab
xPC
AD
DA
Amplifier
Sensors
Valves
Parallel Robot
Hooke joint
Moving
platform
Force sensor
Hydraulic
actuator
Fig. 12. The experimental setup.
Fig. 13. The position tracking errors with PID controller for the trajectories
(x 0:1cospt=8 andy 0:1sinpt=8).
Fig. 14. The position tracking errors with NCFS for the trajectories (x 0:1cospt=8
and y 0:1sinpt=8).
Fig. 15. The orientation tracking errors with PID controller for the trajectories
(c
x
2sinpt=8degrees).
Y. Pi, X. Wang / Control Engineering Practice 19 (2011) 185193 191
performance of the proposed control scheme is similar with that in
the literature (Guo et al., 2008).
6. Conclusions
This study addresses the trajectory tracking control of a 6-DOF
hydraulic parallel robot manipulator with uncertain load distur-
bances. Loaddisturbances are the mainexternal disturbances of the
parallel robot manipulators, and always impact on system perfor-
mance greatly. It is difcult to simultaneously guarantee the
accuracy and robustness for a controller, when the system load
is uncertain and unknown. While most of the reported works
always assume that the load uncertainties are bounded, a novel
control scheme is proposed to deal with uncertain load distur-
bances. Load disturbances are detected by force sensors. Then, a
conventional method is used to design the sliding mode controller
and adaptation laws. The simulation and experimental results
show that the proposed nonlinear control strategy gives a better
performance for the specied tracking task in the presence of
uncertain load disturbances. In this paper, the servo valve model
is crude, and a more evolved model should be considered in
future work.
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
Baron, L., & Angeles, J. (2000). The direct kinematics of parallel manipulators under
joint-sensor redundancy. IEEE Transactions on Robotics and Automation, 16(1),
1219.
Bonchis, A., Corke, P. I., Rye, D. C., & Ha, Q. P. (2001). Variable structure methods in
hydraulic servo systems control. Automatica, 37(4), 589595.
Bonev, I. A., & Ryu, J. (2000). A new method for solving the direct kinematics of
general 66 Stewart platforms using three linear extra sensors. Mechanism and
Machine Theory, 35(3), 423436.
Chen, H. M., Renn, J. C., &Su, J. P. (2005). Sliding mode control with varying boundary
layers for an electro-hydraulic position servo system. International Journal of
Advanced Manufacturing Technology, 26(1-2), 117123.
Dasgupta, B., & Mruthyunjaya, T. S. (2000). The Stewart platform manipulator:
A review. Mechanism and Machine Theory, 35(1), 1540.
Davliakos, I., & Papadopoulos, E. (2008). Model-based control of a 6-dof electro-
hydraulic StewartGough platform. Mechanism and Machine Theory, 43(11),
13851400.
Eryilmaz, B., &Wilson, B. H. (2001). Improved tracking control of hydraulic systems.
Journal of Dynamic Systems Measurement and ControlTransactions of the ASME,
123(3), 457462.
Farhad Aghili, & Mehrzad Namvar (2005). Adaptive control of manipulators using
uncalibrated joint-torque sensing. In Proceedings of the 2005 IEEE international
conference on robotics and automation (pp. 16941699). Barcelona, Spain.
Ferreira, J., De Almeida, F., & Quintas, M. (2002). Semi-empirical model for a
hydraulic servo-solenoid valve. Proceedings of the Institution of Mechanical
Engineers, Part I: Journal of Systems and Control Engineering, 216, 237248.
Gao, X. S., Lei, D. L., Liao, Q. Z., & Zhang, G. F. (2005). Generalized StewartGough
platforms and their direct kinematics. IEEE Transactions on Robotics, 21(2),
141151.
Ghobakhloo, A., Eghtesad, M., & Azadi, M. (2006). Position control of a Stewart
Gough platform using inverse dynamics method with full dynamics. In 9th IEEE
international workshop on advanced motion control (pp. 5055). Istanbul, Turkey.
Guan, C., & Pan, S. X. (2008). Adaptive sliding mode control of electro-hydraulic
system with nonlinear unknown parameters. Control Engineering Practice,
16(11), 12751284.
Guo, H. B., & Li, H. R. (2006). Dynamic analysis and simulation of a six degree of
freedomStewart platformmanipulator. Proceedings of the Institution of Mechan-
ical Engineers Part CJournal of Mechanical Engineering Science, 220(1), 6172.
Guo, H. B., Liu, Y. G., Liu, G. R., & Li, H. R. (2008). Cascade control of a hydraulically
driven 6-DOF parallel robot manipulator based on a sliding mode. Control
Engineering Practice, 16(9), 10551068.
Kim, D. H., Kang, J. Y., & Lee, K. I. (2000). Robust tracking control design for a 6 DOF
parallel manipulator. Journal of Robotic Systems, 17(10), 527547.
Lebret, G., Liu, K., & Lewis, F. L. (1993). Dynamic analysis and control of a Stewart
platform manipulator. Journal of Robotic Systems, 10(5), 629655.
Lee, S. H., Song, J. B., Choi, W. C., & Hong, D. H. (2003). Position control of a Stewart
platform using inverse dynamics control with approximate dynamics. Mecha-
tronics, 13(6), 605619.
Merlet, J. P. (2004). Solving the forward kinematics of a Gough-type parallel
manipulator with interval analysis. International Journal of Robotics Research,
23(3), 221235.
Merritt, H. (1967). Hydraulic control systems. Wiley.
Sirouspour, M. R., & Salcudean, S. E. (2001). Nonlinear control of hydraulic robots.
IEEE Transactions on Robotics and Automation, 17(2), 173182.
Slotine, J., & Li, W. (1991). Applied nonlinear control. Prentice Hall.
Sohl, G. A., & Bobrow, J. E. (1999). Experiments and simulations on the nonlinear
control of a hydraulic servosystem. IEEE Transactions on Control Systems
Technology, 7(2), 238247.
Fig. 16. The orientation tracking errors with NCFS for the trajectories
(c
x
2sinpt=8degrees).
Fig. 17. The force disturbances of each joint actuator (c
x
2sinpt=8degrees).
Fig. 18. The orientation tracking errors with NCFS for the trajectories
(c
x
0:5sin2ptdegrees).
Y. Pi, X. Wang / Control Engineering Practice 19 (2011) 185193 192
Stewart, D. (1965). A platform with six degrees of freedom. Proceedings of the
Institution of Mechanical Engineers, Part A: Journal of Power and Energy, 180(15),
371385.
Tafazoli, S., de Silva C, W., & Lawrence, P. D. (1998). Tracking control of an
electrohydraulic manipulator in the presence of friction. IEEE Transactions on
Control Systems Technology, 6(3), 401411.
Taghirad, H. D., & Nahon, M. (2008). Kinematic analysis of a macromicro redundantly
actuated parallel manipulator. Advanced Robotics, 22(67), 657687.
Ting, Y., Chen, Y. S., & Jar, H. C. (2004). Modeling and control for a GoughStewart
platform CNC machine. Journal of Robotic Systems, 21(11), 609623.
Yao, B., Bu, F. P., Reedy, J., & Chiu, G. T. C. (2000). Adaptive robust motion control of
single-rod hydraulic actuators: Theory and experiments. IEEE/ASME Transac-
tions on Mechatronics, 5(1), 7991.
Zhu, W. H., & Piedboeuf, J. C. (2005). Adaptive output force tracking control of
hydraulic cylinders with applications to robot manipulators. Journal of Dynamic
Systems Measurement and ControlTransactions of the ASME, 127(2), 206217.
Y. Pi, X. Wang / Control Engineering Practice 19 (2011) 185193 193

You might also like