You are on page 1of 8

A SERVO-PNEUMATIC POSITIONING SYSTEM DRIVEN BY FAST SWITCHING ON/OFF VALVES

Cristiano Cardoso Locateli


LASHIP Lab. of Hydraulic and Pneumatic Systems
Mechanical Engineering Department
Federal University of Santa Catarina
Florianpolis, SC, 88040-900
Brazil
cris_locateli@yahoo.com.br
Victor Juliano De Negri
LASHIP Lab. of Hydraulic and Pneumatic Systems
Mechanical Engineering Department
Federal University of Santa Catarina
Florianpolis, SC, 88040-900
Brazil
victor@emc.ufsc.br

Edson Roberto De Pieri
Automation and Control Department
Federal University of Santa Catarina
Florianpolis, SC, 88040-900
Brazil
edson@das.ufsc.br


ABSTRACT
This paper discusses the theoretical and experimental results
of a position control system using a pneumatic actuator driven by
fast switching on/off valves. To begin with, there is a brief
introduction about servo-pneumatic systems driven by fast
switching valves. Subsequently, the mathematical model of the
servo-pneumatic system is discussed involving directional fast
switching on/off valves and the flow control valves, modeled
according to the mass flow equation based on ISO 6358 and the
actuator modeled by Newtons second law and the continuity
equation. The system control is performed through the pulse width
modulation (PWM) technique used with a proportional-
integrative-derivative (PID) controller. The system performance is
analyzed in relation to its application for power control of wind
turbines. The simulations of the servo-pneumatic system were
carried out by Matlab/Simulink with the experimental results
obtained through a test bench.
NOMENCLATURE
A annular area [m
2
]
A
A
annular area of chamber A [m
2
]

A
B
annular area of chamber B [m
2
]
b critical pressure ratio
C sonic conductance [m
5
/N.s]
F
f
friction force [N]
F
c
load force [N]
L actuator stroke [m]
M mass of the actuator piston [kg]
p
1
upstream pressure at the valve [Pa]
p
2
downstream pressure at the valve [Pa]
p
A
pressure in chamber A [Pa]
p
B
pressure in chamber B [Pa]
p
A1
pressure in pipe A [Pa]
p
B1
pressure in pipe B [Pa]
p
s
supply pressure [Pa]
q
m
mass flow rate [kg/s]
q
mA
mass flow rate in chamber A [kg/s]
q
mB
mass flow rate in chamber B [kg/s]
q
mA1
mass flow rate in pipe A [kg/s]
q
mB1
mass flow rate in pipe B [kg/s]
R gas constant [288 J/kg.K]
T
0
temperature at the STP
1
[K]
T
1
temperature at valve inlet [K]
T
A
temperature in chamber A [K]

T
B
temperature in chamber B [K]

1
STP = Standard Condition for Temperature and Pressure adopted by ISO 6358
u control signal
u
A
control signal at valve A
u
B
control signal at valve B
u
dc_A
duty cycle at valve A
u
dc_B
duty cycle at valve B
V
A
volume of chamber A [m
3
]
V
B
volume of chamber B [m
3
]
V
A0
dead volume of chamber A [m
3
]

V
B0
dead volume of chamber B [m
3
]
V
manA
volume of pipe A [m
3
]

V
manB
volume of pipe B [m
3
]
x actuator piston displacement [m]
x& actuator piston speed [m/s]
x& & actuator piston acceleration [m/s
2
]
specific heat ratio

0
density at the STP
1
[kg/m
3
]
INTRODUCTION
The pneumatic position control systems have several
advantages as low cost, high power to weight ratio, easy
maintenance, and environmentally friendly mainly when compared
with hydraulic systems using mineral oil. However some
drawbacks of the pneumatic systems make their control difficult,
such as the air compressibility, nonlinear flow rate through the
valve, nonlinear actuator friction, parameter uncertainties due to
temperature variations, and high sensibility to external
disturbance.
As discussed in [1, 2], position control using pneumatic
systems for speed governors of small hydroelectric power plants
meet the performance requirements of international standards. In
the same way, it is expected that this technology can be used for
power control of wind turbines up to 65 kW, due to its good
performance and low cost. In both types of power plants, the
required forces are compatible with the load capacity of standard
pneumatic cylinders.
However, the proportional directional valve contributes so
significantly to the equipment cost that alternative solutions, such
as fast switching on/off valves, are worth investigations. Their
price can be as low as 10% of the proportional valves and as
presented in this study, the system position error is still minimal
and satisfactory for most applications.
According to [3], the conventional solenoid on/off valves
present some drawbacks regarding the electric and thermal effects,
inertial forces, and friction. However, it seems that the fast
switching solenoid valves give better possibilities for control of
pneumatic systems due to the development of new materials which
allow lower friction and heating and higher valve durability [3].
The mathematical modeling of fast switching on/off valves is
discussed in [4] and [5] considering three subsystems:
electromagnetic, mechanical, and fluid. [4] shows a relationship
between the duty cycle and the average displacement of spool
where it was observed that the transient state behavior can be
neglected due to the fast switching of the valve. Regarding to the
fluid subsystem, there are several ways to model the mass flow
through the valve orifice, either using approximations by
analytical or non-analytical methods [6].
The PWM technique with the duty cycle set by a controller is
one of the most employed methods for pneumatic position control
systems driven by on/off valves. Early work applying the PWM
technique for pneumatic systems dates back to 1987 by Noritsugu
[7, 8] who analyzed the speed control and position control. [9]
shows a comparison between four algorithms for duty cycle
control on the PWM technique. [10] shows a control algorithm
which restricts the maximum and minimum duty cycle of the
valves. [11] presents a theoretical and experimental study of the
dynamics of a servo-pneumatic system driven by fast switching
valves with the PWM technique.
In [12] it is proposed both a MPWM (modified pulse width
modulation) on/off valve control scheme which increased
significantly the performance of the system and a new switching
algorithm for control parameters using a LVQNN (learning vector
quantization neural network) which classifies the external load of
the pneumatic actuator. [13] proposes three linearization
approaches for servo-pneumatic systems driven by PWM signal
and good results were found using a simple linear control strategy.
The use of the nonlinear control (sliding mode) associated with
the PWM technique for control of a servo-pneumatic system is
presented by [14].
Some alternatives to the PWM technique have been proposed
regarding to the position control using on/off valves. [15] applies
the sliding mode control for a pneumatic system where there are
three operation modes that are defined from filling and emptying
of the actuator chamber. [16] shows a sliding mode control with
reduction order where there is only speed feedback. [17] presents
a system control based on the backstepping technique.
As mentioned above, the position control systems based on
on/off valves have previously been studied by several authors.
This paper contributes to this area by presenting a detailed
modeling of the complete system; including valves, actuator, and
controller. Furthermore, the studied system is designed for very
high loads, higher than 4500 N.
Firstly the mathematical system model is detailed, including
the on/off pneumatic valves and the PWM technique with a classic
PID controller. Secondly, the system performance is analyzed
based on theoretical and experimental results including the
pressure behavior in the cylinder chambers. The main conclusions
are presented in the last section.
MATHEMATICAL MODELING
The servo-pneumatic system shown in Fig. 1 comprises two
3/2 fast switching on/off valves driven by solenoid and spring
return (Festo - MHE3-MS1H-3/2G-QS-6) (1V1 (Valve A) and
1V2 (Valve B)), a symmetric double-acting cylinder (Festo -
DNG-125-160-PPV-A-S2) (1A1), a position transducer (Balluff -
BLT5-A11-M0500-P-S32) (1S3), a compressed air source (1Z1),
an air reservoir (1Z3), a pressure reducing valve (1Z2), a filter
(1Z4), two pressure transducers (HBM - HDM P8AP) (1S1 and
1S2), two flow control valves (Camozzi - RFU 483-1/8) (1V3 and
1V4), and a valve controller (Z1).

FIGURE 1. PNEUMATIC POSITIONING SYSTEM
The mass flow rates through the on/off valves are modeled
applying Eq. (1) and (2) those parameters are obtained according
to ISO 6358 [18]. These equations represent both sonic and
subsonic flows.
( ) a w
T
T
p C u q
m
1
0
0 1 0
=

(1)
( )

> |

\
|

<
=
b a for
b
b a
b a for
a w
2
1
1
1

(2)
where
1
2
p
p
a = and u
0
is substituted by u
A
or u
B
when modeling
the on/off valve A or B.
The actuator is modeled by both the continuity equation for
compressible flow and the motion equation. The pressure behavior
in each chamber is given by Eq. (3) with the hypothesis of
isentropic process and absence of external or internal leakage.
( )
( )
mB
B
B
B
B
mA
A
A
A
A
A
q
V x L A
T R
p
V x L A
x A
dt
dpB
q
V Ax
T R
p
V Ax
x A
dt
dp
0
0
0
0
+

+
=
+
+
+
=


&
&

(3)
Since the actuator is symmetrical, A
A
=A
B
=A. The mass flow
rates through the flow control valves (1V3 and 1V4) including the
flow path through the check valves are modeled applying Eq. (1)
and (2) but without the control signal u
0
. The flow control valves
are responsible for both limiting the forward and retracting times
of the actuator and reducing the system vibrations due to
restraining the pressure peaks in the volumes between the on/off
valves and the flow control valves.
The continuity equation applied to these intermediate volumes
(between the on/off valve and the flow control valve) results in
Eq. (4).
( )
( )
1
1
1
1
mB mB
manB
B B
mA mA
manA
A A
q q
V
T R
dt
dp
q q
V
T R
dt
dp
=
=


(4)
Appling Newton's second law, Eq. (5) describes the actuator
movement.
( )
f c B A
F F A p p x M = & &

(5)
The friction force (F
f
=f
V
) is modeled using the variable
viscous friction coefficient (f
V
) as shown in Fig. 2 [19, 20]. The
trajectories "B" (slip) and "C/D" (stick) represent friction forces
for speeds below the speed limit and the curve "A" represents the
friction forces for speeds above the speed limit. The trajectory B
is used when the applied force is greater than static friction force
and the trajectory C/D is used when the applied force is smaller
than static friction force.
Figure 3 shows the experimental static friction map related to
the actuator under study and an approximation through second
order polynomial that is represented by curve A for both
directions of movement. The points of the graph at lower velocity
(negative and positive) sets the value of limit speeds (
Limn
and

Limp
) and the static friction forces (F
Sn
and F
Sp
). The stick speed
(
0n
and
0p
) (trajectory C) are defined as 5% of the limit speed.

FIGURE 2: TRAJECTORY OF THE FRICTION MODEL [20]
-0.04 -0.03 -0.02 -0.01 0 0.01 0.02 0.03 0.04
-400
-300
-200
-100
0
100
200
300
400
Speed (m/s)
F
r
i
c
t
i
o
n

f
o
r
c
e

(
N
)

FIGURE 3. STATIC FRICTION MAP
PWM TECHNIQUE ASSOCIATED WITH PID CONTROL
The position control based on the PWM technique uses a
particular switching frequency where the duty cycle is changed
according to the controller signal. Frequencies between 10 to 60
Hz are being used by several authors [3-13].
The control structure of the servo-pneumatic system is shown
in Fig. 4. The position error is treated by a classic PID controller
resulting in the control signal u. This control signal is changed for
the duty cicle for valve 1 (u
dc_A
) and valve 2 (u
dc_B
) according to
the law described in Eq. (6). This law maintain at least one
chamber being pressurized all times such that is expected to
reduce the system air consumption.
d
y
y
PID
A
u
PWM generator A
PWM generator B

B
u
A
u

B
u
Control
Algorithm
PWM
+

u
A dc
u
_
B dc
u
_
A dc
u
_
B dc
u
_

FIGURE 4. SCHEME OF CONTROL SIGNAL

=
= <
=
= >
1
1 0
1
1 0
_
_
_
_
B dc
A dc
B dc
A dc
u
u u then u if
u u
u then u if

(6)
The u
A
(Fig. 4) is the control signal of valve A that is switched
according both the pulse frequency and duty cycle (u
dc_A
). In the
same way, u
B
is the control signal of valve B based on u
dc_B
. Using
the algorithm, which aims to maintain at least one chamber being
pressurized at all times, it is expected to reduce the system air
consumption.
THEORETICAL AND EXPERIMENTAL RESULTS
The simulations were performed using Simulink/Matlab. A
sinusoidal trajectory with amplitude of 8 mm at a frequency of
0.05 Hz and step inputs of 10 mm and 2 mm were used. These
input signals were chosen for further applications for speed
governors of small hydroelectric power plants and collective pitch
drives for power control of small wind turbines [1, 21-23].
The coefficients C and b of the fast switching on/off valves
and the flow control valves were obtained by tests according to
ISO 6358 standard [18]. Table 1 shows the coefficient values
whereas the other parameters used for simulation can be seen in
Annex A. The flow control valves were adjusted such that the
extending and retracting displacements of the actuator occurs in
five seconds. Annex B shows the block diagram used for the
simulations.
TABLE 1. VALVE COEFFICIENTS
Valve Flow paths C [m
5
/N.s] b
Fast switching valves
1
1-2 6.90x10
-9
0.33
2-3 6.17 x10
-9
0.36
2
1-2 7.03 x10
-9
0.30
2-3 6.16 x10
-9
0.33
Flow control valve
1
1-2 2.80x10
-9
0.56
2-1 5.70x10
-9
0.26
2
1-2 2.74x10
-9
0.60
2-1 4.91x10
-9
0.23

The experiments were carried out on a test bench as shown in
Fig. 5. A dSPACE system was used for data acquisition and
control working together with the Simulink/Matlab and
ControlDesk software. For both the simulations and the
experiments, an integration step of 5x10
-5
s and PID controller
gains of P=1000, I=70, and D=1 were used. A period of 0.033 s
for the PWM signal was used.
Due to the considerable electromagnetic noise in the position
signal, a second-order low pass digital filter with a cut-off
frequency of 70 Hz was used. For the pressure signals first-order
filters of 100 Hz were used.

FIGURE 5. TEST BENCH
Figure 6 shows the comparison between the experimental and
theoretical responses for step inputs of 10 mm and 2 mm.
2 4 6 8 10 12 14
0.072
0.074
0.076
0.078
0.08
0.07
D
i
s
p
l
a
c
e
m
e
n
t

(
m
)
Time (s)


Experimental
Simulation
Input

FIGURE 6. STEP RESPONSES
Both the experimental and simulated responses exhibit similar
behavior with a minor displacement delay. Figure 7 and Fig. 8
show the corresponding position errors that are lower than 0.3 mm
for the steady state at both experimental and simulated responses.
2 4 6 8 10 12 14
-0.01
-0.005
0
0.005
0.01
E
r
r
o
r

(
m
)
Time (s)


Experimental
Simulation

FIGURE 7. POSITION ERRORS
2 4 6 8 10 12 14
-5
-4
-3
-2
-1
0
1
2
3
4
5
x 10
-4
E
r
r
o
r

(
m
)
Time (s)


Experimental
Simulation

FIGURE 8. ZOOM OF POSITION ERRORS
Figure 9 shows the theoretical and experimental pressure
behavior in chambers A and B related to the step inputs shown in
Fig. 6. The theoretical and experimental pressures present similar
behavior.
2 4 6 8 10 12 14
6.2
6.4
6.6
6.8
7
7.2
P
r
e
s
s
u
r
e

(
b
a
r
)
Time (s)
A)


Pa
Pb
2 4 6 8 10 12 14
6.2
6.4
6.6
6.8
7
7.2
P
r
e
s
s
u
r
e

(
b
a
r
)
Time (s)
B)


Pa
Pb

FIGURE 9. PRESSURE BEHAVIOR: A) EXPERIMENTAL;
B) SIMULATION
Considering the pressure response at 3 s, one can observe that
the experimental pressure p
A
suddenly increases while pressure p
B

decreases while the theoretical pressure p
A
do not increase since
that is was already at the supply pressure value. In both theoretical
and experimental results the pressure difference increases until the
piston movement starts, i.e., when the static friction is overcome.
Subsequently, p
A
and p
B
present a constant decay rate until the
steady state position is reached. Figure 10 shows the control signal
(u) regarding these step responses.
2 4 6 8 10 12 14
-1
-0.8
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
0.8
1
C
o
n
t
r
o
l

s
i
g
n
a
l

(
u
)
Time (s)


Experimental
Simulation

FIGURE 10. CONTROL SIGNAL FOR THE STEP INPUTS
Based on the analysis of the step responses, errors, and
pressure behavior s, it is possible to get some conclusions about
the on/off valve control strategy. Analyzing the experimental
responses from 2 s to 3 s, one may notice a position error smaller
than zero. This error causes a control signal of approximately -0.2
as shown in Fig. 10. Consequently, the duty cycle of the valve B
control signal is set at 100% and approximately at 80% for valve
A. Therefore, a pressure difference is produced between the
actuator chambers but it is not enough to overcome the static
friction. The same analysis can be done for the theoretical
responses.
Figure 11 shows the comparison between experimental and
simulated results for a sinusoidal input. Both responses
demonstrate a good trajectory tracking but the real system presents
small steps over the sinusoidal trajectory. These small steps are
mainly caused by the static friction at the actuator and there are
also influences of the flow control valve. Figure 12 shows the
trajectory tracking error.
0 2 4 6 8 10 12 14 16 18 20
0.09
0.092
0.094
0.096
0.098
0.1
0.102
0.104
0.106
0.108
0.11
D
i
s
p
l
a
c
e
m
e
n
t

(
m
)
Time (s)


Experimental
Simulation
Input

FIGURE 11. SINUSOIDAL TRAJECTORY TRACKING
0 2 4 6 8 10 12 14 16 18 20
-8
-6
-4
-2
0
2
4
6
8
10
x 10
-4
E
r
r
o
r

(
m
)
Time (s)


Experimental
Simulation

FIGURE 12. TRAJECTORY TRACKING ERROR
The trajectory tracking errors were as low as 0.7 mm for the
experimental responses and 0.4 mm for the theoretical ones.
Figure 13 shows the pressure behavior in chambers A and B.
0 2 4 6 8 10 12 14 16 18 20
6.6
6.8
7
7.2
P
r
e
s
s
u
r
e

(
b
a
r
)
Time (s)
A)


Pa
Pb
0 2 4 6 8 10 12 14 16 18 20
6.6
6.8
7
7.2
P
r
e
s
s
u
r
e

(
b
a
r
)
Time (s)
B)


Pa
Pb

FIGURE 13. PRESSURE BEHAVIOR FOR SINUSOIDAL
INPUT: A) EXPERIMENTAL; B) THEORETICAL
The divergence between the theoretical and experimental
results is caused by the error in estimating the limit speeds and
static frictions from experiments as shown in Fig. 3. At low
frequency signals as in Fig. 11 the piston movement is under the
stick-slip condition where those parameters have strong influence.
CONCLUSIONS
In this paper, the experimental and theoretical results of a
servo-pneumatic system driven by fast switching on/off valves for
step and sinusoidal inputs were discussed. The on/off and flow
control valves were modeled by the compressible flow equation
using experimental parameters obtained according to ISO 6358
standard. Newtons second law, including the friction force
modeled by the variable viscous friction coefficient, and the
continuity equation were used to describe the pneumatic actuator
and pipe behavior. The PWM technique with a PID classic control
was also used.
For both step and sinusoidal inputs there were low position
errors demonstrating that the fast switching on/off valves can be a
promising solution considering the cost benefit when compared to
proportional directional valves. The control strategy adopted in
this research, where one valve is completely open while the other
is being controlled, reduces both the air consumption and the
number of cycles of valve operation, when compared with the
control methods most employed.
The authors are implementing a variable structure control
with sliding modes aiming to compare with the results presented in
this paper.
ACKNOWLEDGMENTS
The authors are grateful to CNPq (Brazilian National Council
for Scientific and Technological Development) for the financial
support.
REFERENCES
[1] Asaff,Y. E., and De Negri, V. J., 2007. Development of a
High Power Pneumatic Servo-Positioning System for Speed
Governors of Hydraulic Turbines. Proceedings of the 19th
International Congress of Mechanical Engineering, COBEM
2007, Brasilia, Brazil.
[2] Asaff, Y. E., Oliveira, L. G., and De Negri, V. J., 2008.
Applicability of Servo-pneumatic Positioning Systems for
High Loads, Bath/ASME Symposium on Fluid Power and
Motion Control (FPMC08), Bath, UK; 219-32.
[3] itum, ., ili, T. and Essert, M., 2007. High speed
solenoid valves in pneumatic servo applications
Mediterranean Conference on Control and Automation, pp.
1-6.
[4] Taghizadeh, M., Ghaffari, A., and Najafi, F., 2009. Modeling
and identification of a solenoid valve for PWM control
applications. C R Mecanique 337, pp. 131140.
[5] Topu, E. E., YukseI, I., and Kamis, Z., 2006. Development
of electro-pneumatic fast switching valve and investigation of
its characteristics Mechatronics 16, pp. 365378.
[6] Jouppila, V., Gadsden, A., and Ellman, A., 2010. Modeling
and Identification of a Pneumatic Muscle Actuator System
Controlled by an On/Off Solenoid Valve. 7th International
Fluid Power Conference. Aachen.
[7] Noritsugu, T., 1987. Development of PWM mode electro-
pneumatic servomechanism, Part I: Speed control of a
pneumatic system. J Fluid Control 17(1):6579.
[8] Noritsugu, T., 1987. Development of PWM mode electro-
pneumatic servomechanism, Part II: Position control of a
pneumatic system. J Fluid Control 17(2):728.
[9] Van Varseveld, R. B., and Bone, G. M., 1997. Accurate
position control of a pneumatic actuator using on/off solenoid
valves, IEEE/ASME Transactions on Mechatronics 2 3, pp.
195204.
[10] Gentile, A., Giannncearo, N., and Reina, G., 2002.
Experimental tests on position control of a pneumatic
actuator using on/off solenoid valves. IEEE lCIT'O2,
Bangkok, THAILAND.
[11] Messina, A., Giannoccaro, N. I., Gentile, A., 2005.
Experimenting and modelling the dynamics of pneumatic
actuators controlled by the pulse width modulation (PWM)
technique. Mechatroincs 15: 859-881.
[12] Ahn, K., and Yokota, S., 2005. Intelligent Switching Control
of Pneumatic Actuator Using On/Off Solenoid Valves.
Mechatronics 15:683-702.
[13] Taghizadeh, M., Ghaffari, A., Najafi, F., 2008. A
linearization approach in control of PWM-driven servo-
pneumatic systems. Proceedings of the 40th IEEE/SSST
Sym., New Orleans, LA, USA. pp. 395399.
[14] Shen X., Zhang, J., Barth, E. J. and Goldfarb, M. 2006,
Nonlinear model-based control of pulse width modulated
pneumatic servo systems, ASME Journal of Dynamic
Systems, Measurement, and Control. Vol. 128, pp. 663669.
[15] Nguyen, T., Leavitt, J., Jabbari, F., and Bobrow, J. E., 2007.
Accurate sliding-mode control of pneumatic systems using
low-cost solenoid valves. IEEE/ASME Transaction
Mechatronics 12, pp. 216219.
[16] Paul, A. K., Mishra, J. K., and Radke, M. G., 1994. Reduced
Order Sliding Mode Control for Pneumatic Actuator, IEEE
Transactions on Control Systems Technology., 230, pp. 271
276.
[17] Langjord, H., and Johansen, T. A., 2010. Dual-mode
switched control of an electropneumatic clutch actuator,
IEEE/ASME Transaction on Mechatronics,
DOI:10.1109/TMECH.2009.2036172.
[18] ISO, 1989. ISO 6358, Pneumatic Fluid Power
Components using Compressible Fluids Determination of
Flow-rate Characteristics.
[19] Gomes, S. C. P.; and Rosa, V. S. 2003. A new approach to
compensate friction in robotic actuators. International
conference on robotics and automation, Taipei, Taiwan.
[20] Machado, C., 2003. Compensao de atrito em atuadores
hidrulicos utilizando redes neurais. 86 p. Master`s Thesis,
Universidade Federal de Santa Catarina, Florianpolis,
Brazil.
[21] IEC 61400-2, Ed. 2.0, 2006. Wind Turbines-Part 2: Design
Requirements for Small Wind Turbines International
Electrotechnical Commission, Geneva.
[22] Bossanyi, E. A., and Jamieson, P., 1999. Blade pitch system
modeling for wind turbines. Proceedings of the 1999
EuropeanWind Energy Conference, Nice, pp. 893896.
[23] Xin,, G., and Jing, H., 2008. "Study on a novel hydraulic
variable-pitch system of wind turbine." Industrial Technology.
ICIT 2008. IEEE International Conference on. :1-4.
ANNEX A
SYSTEM PARAMETERS
TABLE 2. PARAMETERS FOR SIMULATION
Smbolo Descrio Valor
T
0
temperature at the STP 293.15 K
T
1
air temperature at valve inlet 293.15 K
R gas constant 288 J/kg.K
T
A
, T
B
temperature in chambers of actuator 293.15 K
V
manA
volume of pipe A 9 x 10
-6
m
3

V
manB
volume of pipe B 9 x 10
-6
m
3

V
A0
dead volume of chamber A 2.83210
-5
m
3

V
B0
dead volume of chamber B 2.8310
-5
m
3
A annular area 1.15 x 10
-2
m
2
L actuator length 0.160 m

0
density at the STP 1.22 kg/m
3

specific heat ratio 1.4
M mass of actuator piston 2.16 kg
p
s
supply pressure 7 bar

The main structure of the Simulink/Matlab diagram used to
validate the position control system is presented in ANNEX B.
ANNEX B
SIMULATION DIAGRAM

You might also like