Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Designing Robust Force Control of

Despite System and Environmental Uncertainties

By Navid Niksefat
and Nariman Sepehri

T
his article presents the de-
sign of a robust force con-
troller for a hydraulic
actuator interacting with an
uncertain environment via
quantitative feedback the-
ory (QFT). After the derivation of a realistic
nonlinear differential equation model, a
linearized plant transfer function is devel-
oped. The effects of nonlinearities are ac-
counted for by describing the linearized
model parameters as structured uncer-
tainty. The impact of environmental vari-
ability as well as variations in hydraulic
component parameters are also included
as uncertainty in the model. The QFT de-
sign procedure is carried out to design a ro-
bust controller that satisfies performance
specifications for tracking and disturbance ©ABB Control Valves.
rejection. The designed controller enjoys
the simplicity of fixed-gain controllers, is
easy to implement, and at the same time is
robust to the variation of hydraulic func-
tions as well as environmental stiffness.
The controller is implemented on an industrial hydraulic ac- Introduction
tuator equipped with a low-cost proportional valve. The ex- Many industrial applications, such as manufacturing auto-
perimental results show that robust stability against system mation and material handling, involve interaction with the
uncertainties and under varying conditions is achieved and environment. In these applications, force control is re-
the performance goals are satisfied. quired. Hydraulic actuators are advantageous for such ap-

Niksefat and Sepehri (nariman@cc.umanitoba.ca) are with the Department of Mechanical Engineering, The University of Manitoba, Win-
nipeg, Manitoba, R3T 5V6 Canada.

0272-1708/01/$10.00©2001IEEE
66 IEEE Control Systems Magazine April 2001
plications because of their high force-to-weight ratio and switching control scheme using a Lyapunov-based adaptive
fast response time. Additionally, they are able to maintain law to reduce parametric uncertainty. The implementation
their loading capacity indefinitely, which would usually of the controller, which is based on the measurements of po-
cause excessive heat generation in electrical components sition, velocity, acceleration, pressure, and spool displace-
[1]. Unlike in electric actuators, however, force control in ment, showed good performance for high-frequency force/
hydraulic actuators is a difficult problem [2], [3]. pressure tracking.
In a hydraulic actuator, the control signal activates the Wu et al. [10] applied a generalized predictive control al-
spool valve that controls the flow of hydraulic fluid into and gorithm to a hydraulic force control system. The controller
out of the actuator. This flow in turn causes a pressure dif- was experimentally evaluated for various environmental
ferential buildup that is proportional to the actuator force. stiffnesses and setpoints. The method, however, relies
Even if the spool valve dynamics are ignored, the control heavily on online parameter estimation and consequently is
signal fundamentally controls the derivative of the actuator computationally expensive.
force and not the force itself. Furthermore, hydraulic sys- Laval et al. [11] used an H ∞ approach to robustly control
tems are highly nonlinear and subject to parameter uncer- the force exerted by a double-acting symmetric hydraulic
tainty; parameters change with time as a result of variations
cylinder with a servovalve. The importance of uncertainties
in operating conditions and component degradation. For ex-
in the environment, measurement, and nonlinearities on the
ample, the supply pressure is subject to variation that may
performance of hydraulic force control systems was high-
be generated by the operation of other actuators in a
lighted. Limited test results, demonstrating the achieve-
multiuser environment [4]. The flow and pressure coeffi-
ment of a stability/performance trade-off utilizing an H ∞
cients, characterizing fluid flow into and out of the valve, are
approach, were presented.
functions of load and supply pressure and can vary under
Despite the existence of a great number of force control
different operating conditions [5]. Also, the effective bulk
concepts, methods, and algorithms, there is still a large gap
modulus in hydraulic systems can significantly change un-
between theory and industrial practice. The reasons have
der various load conditions, oil temperature, and air con-
tent in the oil [6]. Design of a controller in the face of such a been ascribed to the poor industrial control architecture,
range of parameter variations and disturbances is challeng- which does not allow the implementation of sophisticated
ing. This article presents the application of QFT to the de- algorithms [12]. In this article, we employ the QFT tech-
sign of a robust force controller for hydraulic actuators. nique to design an explicit force controller for an industrial
In the literature, several force control strategies have hydraulic actuator. The goal is to arrive at a fixed-gain con-
been proposed for hydraulic actuators. Conrad and Jensen troller that: 1) is of low order and easy to implement, 2) is ro-
[2] used combinations of velocity feedforward, output feed- bust against uncertainties in both environmental stiffness
back, and a Luenberger observer with state estimate feed- and actuator functions, and 3) does not require exact knowl-
back for force control of a double-rod hydraulic actuator. edge of the system’s parameters.
The simulation and experimental results for a constant QFT is a robust controller design methodology aimed at
setpoint force showed superior performance of the pro- plants with parametric and unstructured uncertainties. The
posed method over conventional (P or PI) force feedback concept was first introduced by Horowitz in the early sixties
controllers. However, the variations of load and supply and was later refined by him and others into a technique
pressure were not considered in their study. [13]-[15]. QFT emphasizes the fact that feedback is only nec-
Chen et al. [7] designed a sliding-mode controller for a essary because of uncertainty and that the amount of feed-
single-rod hydraulic actuator interacting with a spring as an back should therefore be directly related to the extent of
environment. Using position, velocity, acceleration, force, plant uncertainty and unknown external disturbances. Mini-
and pressure feedback, the variable-structure controller mizing the cost of feedback, as measured by the amount of
proved to be capable in both static and dynamic force con- controller bandwidth, is the main objective of QFT [16].
trol tasks. The effect of servo-amplifier gain variation was Therefore, the plant uncertainty and the closed-loop toler-
also examined; however, the effect of variations in environ- ances are formulated quantitatively so that the cost of feed-
mental stiffness was not studied. back can be assessed at each stage of the design process
Sun et al. [8] employed a sliding-mode controller with a [17]. (See the sidebar “An Overview of QFT” on p. 68.) The
perturbation observer for a single-rod electrohydraulic sys- method has been applied to a wide range of engineering
tem. The effect of cylinder position and velocity on the pres- problems, including flight control [18], robot position con-
sure dynamics was considered as perturbation to the trol [19], and manufacturing systems [20]. Regarding the ap-
control model, which was estimated by an observer. The ex- plication of QFT to hydraulic systems, the technique was
perimental results verified improved steady-state and tran- used to control an electrohydrostatic actuator as part of a
sient performance as compared with traditional flight control system by Pachter et al. [21]. The controller’s
proportional-integral-differential (PID) controllers. efficacy was validated by simulation results. Thompson and
Adaptive control strategies have also been considered Kremer [22] developed a QFT controller for a variable-dis-
for hydraulic force control. Liu and Alleyne [9] developed a placement pump system based on a linearized model with

April 2001 IEEE Control Systems Magazine 67


An Overview of QFT

Q uantitative feedback theory (QFT) is a frequency domain method


intended for practical control system design given robust perfor-
mance specifications. Its goal is to design a low-bandwidth controller
α 0 , yields the nominal plant P0 ( s ) = P ( s , α 0 ) . Many problems of
practical interest can be expressed by the above plant model with
parametric uncertainty.
that satisfies performance specifications, despite system uncertainties
QFT Design Problem
and disturbances. A low-bandwidth controller is a key issue in any
practical design to avoid problems with noise amplification, resonance, The QFT design problem is to synthesize a pair of strictly proper,
and unmodeled dynamics. A brief outline of the classical QFT problem rational, stable functions G( s ) and F ( s ) such that the following
is presented here. A full treatment is given in [13] and [15]. Short over- specifications are satisfied while the bandwidth of the controller
views can also be found in [28] and [29]. is kept as low as possible.

QFT Configuration Robust Stability: The closed-loop system

A two degree-of-freedom system is typically assumed for the QFT L( s ,α )


technique when only output Y and command input R can be T ( s ,α ) =
1 + L( s ,α )
measured independently. Plant P contains system uncertainties
and is exposed to unknown disturbance D. G is the controller, must be stable ∀ α ∈ Ω. L( s , α ) = P ( s , α ) G( s ) is the open-loop
which can help reduce the variation of the plant output due to transfer function.
uncertainties and disturbances, while F is used merely as a filter to
Robust Tracking: A tracking specification for the reference
tailor the response to meet the control system’ s specifications.
input is given as
The uncertain plant can be described as
m
Tl (ω ) ≤ |F (iω )T (iω ,α )| ≤ Tu (ω ) ω ∈ [0 , ∞ )
∑ p (α )s
i
i

P( s ,α ) = i=0
n
where the frequency response bounds, Tl (ω ) and Tu (ω ) , are
∑ q (α )s
i=0
i
i

specified by the designer.

where α ∈ Ω ⊂ R p is an uncertain parameter vector. Ω is a Disturbance Attenuation: The requirement for disturbance
compact set of parameter variations and may be given as rejection at plant output is expressed as

Ω = {α :α i ∈ [α i ,α i ] ,i = 1,… , p} max TD ( jω ,α ) ≤ M D (ω )
α ∈Ω

where each uncertain parameter α i ( i = 1,... , p ) varies indepen-


dently within an interval. Choosing a reference parameter vector, where

parametric uncertainty. The closed-loop frequency domain Electrohydraulic Actuator Modeling


specifications were satisfied, and the corresponding step
responses were found to be reasonable. The successful ap- Presentation of
plication of this method to a wide range of engineering prob- Nonlinear Dynamic Equations
lems provides evidence of its potential effectiveness for A schematic diagram of a hydraulic actuator in contact
with the environment is shown in Fig. 1. In this study, the
hydraulic systems. One objective of this study is to evaluate
well-known model of manipulator-sensor-environment
and demonstrate the benefits of QFT when applied to hy-
[23], [24] is coupled with the nonlinear hydraulic actua-
draulic force control systems.
tor dynamics. The sensor with stiffness ks and damping ds
The article is organized as follows. In the next section, we
connects the actuator piston, represented by the mass
derive a mathematical model of a hydraulic actuator in con-
ma , to the environment. The environment is represented
tact with the environment. This model is then used to con- by mass me , damping de , and stiffness ke . The dynamic
struct a parametrically uncertain linear transfer function to equations are
describe the system. The uncertainty reflects variation in
operating-point-dependent parameters, changes in the en- ma x = fa − ds ( x − x e ) − dx − ks ( x − x e ) (1)
vironment, and changes in hydraulic actuator’s functions.
Then, we use the QFT method to design a single fixed-gain
robust force controller for the above uncertain plant. Fol- me xe = ds ( x − x e ) − de x e + ks ( x − x e ) − ke x e (2)
lowing that, we demonstrate the controller’s performance
on an experimental hydraulic actuator test rig. Finally, we
present conclusions. f = ks ( x − x e ) (3)

68 IEEE Control Systems Magazine April 2001


1 frequency to keep the controller bandwidth small. During this
TD ( s ,α ) = stage, the designer should effect a trade-off between conflicting
1 + L( s ,α )
specifications, controller complexity, and the cost of the
feedback in the bandwidth. Once a satisfactory L0 ( s ) is arrived
is the transfer function from the disturbance to the output and at, the controller can be extracted from L0 ( s ) by dividing it by the
MD (ω ) is the magnitude of disturbance rejection. nominal plant transfer function P0 ( s ).
QFT Design Procedure 4) Design of prefilter: Design of a proper L0 ( s ) only guaran-
The design procedure involves the following steps. tees that the variation in the closed-loop transfer function,
T ( s , α ) , is less than or equal to that allowed. Therefore, a
1) Generating the plant templates: The region in the Nichols prefilter is required to bring the response within the upper
chart occupied by the complex values of P ( jω , α ) at each and lower tolerances, Tl (ω ) and Tu (ω ) . The prefilter magni-
frequency is called the plant template (see the figure at right). tude bound at each frequency can be calculated from the ro-
This essentially characterizes the plant uncertainty by capturing bust tracking specification.
the gain and phase variations of the plant at a given frequency.
The templates are used to generate bounds on the Nichols chart
for the controller design.

2) Generating performance bounds: The design specifications


60 L0(jω1)
outlined above should be translated into certain bounds on the ω1
50
nominal open-loop transfer function, L0 ( s ) = P0 ( s ) G( s ) , to Plant Template at ω = ωi
40
reveal the trade-offs between performance specifications and

Open-Loop Gain (dB)


30 ω2
robustness at each frequency. These bounds are derived either
20
by moving the plant template between the closed-loop
10 QFT Bounds ω3
magnitude contours on the Nichols chart or by a computer
search. Examples of bounds are shown in the figure. These 0
−10 ω4
bounds are used in the next step as a guide for loop shaping the
nominal open-loop transfer function. −20 Nominal Loop, L0(jω)
−30 ω4
3) Loop shaping: Once the QFT bounds are determined, the
−40
nominal loop transfer function, L0 ( s ), should be designed, by −350 −300 −250 −200 −150 −100 −50 0
Open-Loop Phase (deg)
adding proper poles and zeroes, to yield a stable nominal closed
loop, while at the same time satisfying all bounds. A typical plot
of L0 ( s ) is given in the figure. An optimum L0 ( s ) is one that
satisfies the bounds and decreases as rapidly as possible with Typical bounds and nominal loop function in the Nichols chart.

where f is the sensed force and x, x e , and d are the actua- 2 2


tor’s displacement, the environmental displacement, and q i = cdwx sp ( pi − pe ) q o = cdwx sp ( ps − po )
ρ ρ (5b)
the equivalent viscous damping coefficient of the actua-
tor, respectively. The force originated from the hydraulic
actuator, fa , is where q i and q o represent fluid flows into and out of the
valve, respectively. cd is the orifice coefficient of dis-
fa = pi Ai − po Ao (4) charge, ρ is the mass density of the fluid, ps is the pump
pressure, and pe is the return (exit) pressure. w is the area
where Ai and Ao are the piston effective areas and pi and po gradient that relates the spool displacement( x sp ) to the or-
are the input and output line pressures, respectively. The ifice area. Continuity equations for oil flow through the cyl-
governing nonlinear equations describing the fluid flow dis- inder, neglecting the leakage flow across the actuator’s
tribution in the valve are written, in their simplest forms, as piston, are
[25]
x sp ≥ 0 (extension) dx 1 dpi
q i = Ai + Vi
dt β dt (6a)
2 2
q i = cdwx sp ( ps − pi ) q o = cdwx sp ( po − pe )
ρ ρ (5a)
dx 1 dpo
q o = Ao − Vo .
x sp < 0 (retraction) dt β dt (6b)

April 2001 IEEE Control Systems Magazine 69


Here β is the effective bulk modulus of the hydraulic fluid, as uncertainty. In this section, the derivation procedure will
and Vi and Vo are the volumes of fluid trapped at the sides of be followed step-by-step to highlight the assumptions that
the actuator. They can be expressed as functions of actua- are made during the derivation and simplification.
tor displacement The linearized flow relationships from (5) are

q i = K si x sp − K pi pi
Hydraulic actuators are (9a)

advantageous because of their


q o = K so x sp + K po po
high force-to-weight ratio and (9b)

fast response time. where K si (K so ) and K pi (K po ) represent flow and pres-


sure sensitivity gains of the valve, respectively. Super-
scripts i and o stand for input and output, respectively.
Vi ( x ) = Vi + xAi (7a) They are given as follows:

for extension ( x sp ≥ 0 ) for retraction ( x sp < 0 )


Vo ( x ) = Vo − xAo (7b) 2 2
K si = cdw ( ps − pi ) K si = cdw ( pi − pe )
ρ ρ
where Vi and Vo are the initial volumes trapped in the blind 2 2
and rod sides of the actuator. The relationship of the spool K so = cdw ( po − pe ) K so = cdw ( ps − po )
ρ ρ
displacement, x sp , and the input voltage, u, to the propor-
cdwx sp − cdwx sp
tional valve can be expressed as a first-order system that is K pi = K pi =
accurate enough to model low-cost proportional valves [26] 2ρ( ps − pi ) 2ρ( pi − pe )
cdwx sp − cdwx sp
K po = K po = .
τ dx sp 1 2ρ( po − pe ) 2ρ( ps − po )
u= + x sp
ksp dt ksp (8)

As can be seen, K si (K so ) and K pi (K po ) are load- and pres-


where τ and ksp are gains describing the valve dynamics.
sure-dependent variables. Hence, in this study, they are
Equations (1)-(8) express the relationship between the con-
considered uncertain but bounded parameters. Assuming
tact force, f, and the input control voltage, u.
small piston displacements within the vicinity of the
midstroke, the following approximation is made:
Derivation of Linear Transfer Function
The nonlinear equations derived above are now linearized
Vi ( x ) Vo ( x ) 1  Vi + Vo 
about an operating point to obtain a linear transfer function. ≈ ≈   = C.
The objective is to utilize this model within the QFT ap- β β β 2  (10)
proach. As will be seen later, the variation of the operating-
point-dependent parameters will be included in the model Thus, (6) can be written in the Laplace domain as

x xe Qi = Ai sX + CsPi (11a)
ks ke
po
pi Ai ma me
Ao Qo = Ao sX − CsPo . (11b)
ds de
qi qo d
Substituting (9) into (11) and rearranging for line
pressures, we have

xsp
Ai s K si
Pi = − X+ X sp
Cs + K p
i
Cs + K pi (12a)

Pe Pe
Ps
Ao s K so
Figure 1. Schematic model of the hydraulic actuator interacting with the Po = X − X sp .
environment. Cs + K po Cs + K po (12b)

70 IEEE Control Systems Magazine April 2001


Incorporating (12) into (4) we have Table 1. Operating values and parameter ranges
pertaining to the linear transfer function.
 KoA K si Ai   Ai2 s Ao2 s  Parameter Nominal Value Range
Fa =  s o o +  X sp −  + o 
X.
Cs + K p Cs + K p  Cs + K p Cs + K p  (13)
i i
ke 75 (kN/m) 50-100
3
Equations (1)-(3) are now combined to form the following Ks 0.375 (m /pa.s) 0.25-0.5
transfer functions: −12 2 −12
Kp 2.5(10) (m /s) 0-5(10)
−11 −11 −11
Fa ( s ) 1 ( ma s + ds )( me s + ( de + ds )s + ks + ke )
3
2 2 C 1.5(10) (m /pa) 1(10) -3(10)
= 
F ( s ) ks  me s 2 + de s + ke d 700 (N/m/s) 600-800

( ds s + ks )( me s 2 + de s + ke )  ma 20 (kg) 19.9-20.1
+ 
me s 2 + de s + ke  Ai 0.00203 (m )
2
0.00193-0.00213
(14)
2
Ao 0.00152 (m ) 0.00144-0.00160

X ( s ) 1 me s 2 + ( de + ds )s + ks + ke ksp 0.0012 (m/V) 0.0011-0.0013


= .
F ( s ) ks me s 2 + de s + ke (15)
τ 35 (ms) 30-40

Substituting Fa ( s ) and X ( s ) from (14) and (15) into (13) and


rearranging it, the transfer function describing the relation Including the dynamics of the valve from (8), the transfer
between contact force and spool displacement is derived as function between measured contact force, F ( s ), and control
(1 6 ) , s h o w n b e l o w , w h e r e ϕ( s ) = me s 2 + de s + ke , voltage, U ( s ), is written as
ψ( s ) = ma s 2 + ds, and ζ( s ) = ds s + ks .
To simplify the above transfer function, K si (K so ) and K pi F (s) ksp  K s ke ( Ai + Ao ) 
o =  .
( K p ) are replaced by K s and K p . Hence transfer function U ( s ) ( τs + 1) ( K p + Cs )( ma s + ds + ke ) + ( Ai s + Ao s ) 
2 2 2

(16) is reduced to (17), shown below.


(20)
Any variations or approximation in K si (K so ) and K pi (K po )
Equation (20) is now considered as a parametrically uncer-
are now included as uncertainty in K s and K p , respectively.
tain system. For example, the uncertainty ranges in K s and
The above relation can be further simplified if the stiffness
of the force sensor and the piston rod are high compared K p reflect variations in the operating point, supply pres-
with the environmental stiffness and the hydraulic compli- sure, and, in part, the orifice area gradient. The variations
ance. Hence their dynamics are not excited during the con- in the environmental stiffness and damping of the system
tact and can be lumped together as a rigid body. Based on are included in parameters ke and d, respectively. Further-
this assumption, (17) is rewritten as more, the uncertainty in C represents the changes in the
fluid bulk modules and the volumes of fluid trapped at the
sides of the actuator. The uncertainty in the valve charac-
F (s) K s ( Ai + Ao )( me s 2 + de s + ke )
= . teristic is captured in variations of τ and ksp . All these pa-
X sp ( s ) ( K p + Cs )( ma s 2 + ( d + de )s + ke ) + ( Ai2 s + Ao2 s )
rameters are known to have an effect on the system’s
(18) stability.
Further, we assume the dynamics of the environment is
Table 1 lists the nominal values of all parameters in (20)
dominated by a pure stiffness, ke . This type of environment
and their corresponding uncertainty ranges. These values,
has enjoyed popularity among many researchers [27]. The
which are representative of the hydraulic actuator under in-
system transfer function is finally written in its simplest
vestigation (see the “Experimental Setup” section), were ob-
form as
tained from available references (including manufacturer
specifications), direct measurements of some geometric pa-
F (s) K s ke ( Ai + Ao )
= . rameters, and indirect identification based on experimental
X sp ( s ) ( K p + Cs )( ma s 2 + ds + ke ) + ( Ai2 s + Ao2 s) (19) observation of the responses [26]. These uncertain parame-

F (s)
=
(
ks ( me s 2 + de s + ke ) K si Ai ( K po + Cs ) + K so Ao ( K pi + Cs ) )
X sp ( s ) ( K pi + Cs )( K po + Cs )( ψ ( s )ϕ( s ) + ζ( s )ϕ( s ) + ψ ( s )ζ( s )) + ( Ai2 s( K po + Cs ) + Ao2 s( K pi + Cs ))(ϕ( s ) + ζ( s )) (16)

F (s) ks K s ( Ai + Ao )( me s + de s + ke )
2

= . (17)
X sp ( s ) ( K p + Cs )(( ψ ( s )ϕ( s ) + ζ( s )ϕ( s ) + ψ ( s )ζ( s )) + ( Ai2 s + Ao2 s )(ϕ( s ) + ζ( s ))

April 2001 IEEE Control Systems Magazine 71


Controller Synthesis
Prefilter Controller Uncertain Plant D (s)
The goal of this section is to design a robust force
controller for the hydraulic actuator that is rep-
R(s)
F (s)
+
G(s) P(s,α) + Y(s)
resented by the uncertain transfer function (21).
− +
A typical two-degree-of-freedom feedback sys-
tem configuration in QFT is shown in Fig. 2. A
strictly proper controller, G( s ), and a strictly
proper prefilter, F ( s ), are to be designed such
that the following conditions are satisfied (read-
Figure 2. Two-degree-of-freedom QFT control system. ers are referred to [28] for a description of the
QFT design procedure):
ters are grouped into a vector, denoted as α. Then the sys- i) Closed-loop robust stability. The associated QFT ro-
tem’s open-loop transfer function can be defined as bustness constraint in terms of the nominal loop
transfer function, L0 ( s ) = P0 ( s )G( s ), is given by [28]:
F (s)
P( s ,α ) = .
U(s) (21)  
 L0 (iω ) 
The plant evaluated at its nominal operating point is re-  p (iω )  ≤ M = 14
. ∀ω ∈ [0 , ∞ )
 0
+ L0 (iω )
ferred to here as the nominal plant and is shown as  p(iω ,α )  (23)

P0 ( s ) = P( s ,α 0 ) (22) which implies an approximately 3 dB gain margin for


the closed-loop system.
where α o is the vector containing all nominal values of the
plant transfer function.

ω=0.01
120
20 100
Upper Bound ω=0.05
0 80 ω=0.1
Magnitude (dB)

−20 60
ω=0.5
Magnitude (dB)

Tu (s)
−40 40 ω=1

−60 Lower Bound 20 ω=5


ω=70 ω=10
−80 Tl (s) 0 ω=100
ω=50
−100 −20
−350 −300 −250 −200 −150 −100 −50 0
−120 Phase (deg)
100 101 102 (a)
Frequency (rad/s)
(a) ω=0.01
120 0.01
1.4
100
1.2 Upper Bound 0.05 0.05
80 0.1
Magnitude (dB)
Force (Normalized)

1
60 0.5
0.8 0.5
Lower Bound 40 1 1
0.6 5
20 10 5
0.4 10
0 50
0.2 100 70
−20
−350 −300 −250 −200 −150 −100 −50 0
0 Phase (deg)
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Time (s) (b)
(b)
Figure 4. QFT bounds on Nichols chart with: (a) uncompensated
Figure 3. Frequency and time domain tracking bounds. plant and (b) nominal loop.

72 IEEE Control Systems Magazine April 2001


ii) Robust reference input. To meet tracking performance transfer function, L0 , should be determined such that it lies
requirements, the controller should satisfy the follow- on or above the corresponding bounds at every frequency.
ing inequality: This can be done using a procedure that is called loop shap-
ing in QFT. Finally, the controller can be extracted from L0 by
dividing by the nominal plant transfer function, P0 .
L0 (iω ) Design frequencies are selected as ω = {0.01,0.05,0.1,
Tl (iω ) ≤ F (iω ) ≤ Tu (iω ) ∀ω ∈ [0, ∞ )
p0 (iω ) 0.5,1,5,10,50,70,100} rad/s. The bounds generated by con-
+ L0 (iω )
p(iω ,α ) (24)
straints (23), (24), and (26) are shown in Fig. 4(a) for the un-
compensated plant. The nominal open-loop transfer func-
where the upper and lower tracking bounds are de-
fined as 20
Upper Bound δ
0
 s 
 + 1
 2.8  −20
Tu ( s ) =

Magnitude (dB)
s s s 
 + 1  + 1  + 1 −40
4 7 8  (25a) δall
Lower Bound
−60

−80
1
Tl ( s ) = .
s 9.6 s 
2
 s  s −100
 + 1  + 1  + + 1
 4.8   80   50 50  (25b) −120
100 101 50 102
Frequency (rad/s)
These bounds are built from the time domain figures of (a)
merit for step responses such as peak overshoot, peak time,
and settling time [14]. In this work, the desired lower track- 20
ing bound, Tl ( s ), is built to have an overdamped response Upper Bound
0
with ≈1 s settling time. For this purpose, a model with a real
pole at s = −4.8 and a pair of complex poles are chosen. The −20
Magnitude (dB)

real pole must be more dominant than the complex poles −40
[14]. Moreover, a high-frequency pole at s = −80 is inserted Lower Bound
in Tl ( s ), which does not affect the desired performance −60
specification but widens the range between Tu ( s ) and Tl ( s ) −80
in the high-frequency band. The figures of merit for upper
bound, Tl ( s ), are a 5% peak overshoot and ≈1 s settling time. −100
Tu ( s ) is selected with three real poles and a zero. The zero is −120
closer to the origin than the poles to have an underdamped 100 101 102
Frequency (rad/s)
response. The frequency and time domain plots of these (b)
bounds are shown in Fig. 3(a) and (b), respectively.
1200
iii) Closed-loop disturbance attenuation (sensitivity reduc- Upper Bound
tion). For disturbance rejection at the plant output, an 1000
upper tolerance is imposed on the sensitivity function.
Here we consider only a constant upper bound to limit 800 Lower Bound
Force (N)

the peak value of disturbance amplification as follows:


600

400
 p0 (iω ) 
 p(iω ,α )  200
S (iω ,α ) max =  ≤ M D (ω ) = 1.2 ∀ω ∈ [0 ,10].
p (iω )
 0 + L0 (iω )
 p(iω ,α ) max 0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
(26) Time (s)
(c)
The above design specifications impose constraints on Figure 5. System responses over range of parametric uncertainty
the allowable loop gain,|L0|. These constraints can be shown (simulation): (a) closed-loop frequency responses without prefilter,
on a Nichols chart as a curve boundary at each design fre- (b) closed-loop frequency responses with prefilter, and (c) time
quency. Once the bounds are derived, the nominal open-loop responses.

April 2001 IEEE Control Systems Magazine 73


tion, L0 , must be shaped to satisfy all these bounds at each The next step is to add two high- frequency poles to allow for
frequency. For low frequencies, L0 should lie on or above the a quick descend of the nominal loop to decrease the cost of
bounds to satisfy the constraints. However, for higher fre- feedback. Fig. 4(b) shows the final loop shaping of the sys-
quencies in the range, e.g., ω = 50, 70, and 100 rad/s, the con- tem. The controller that satisfies the specifications is
straints generate closed boundaries which L0 should not
enter. For the industrial hydraulic actuator under investiga- s  s 
0.0065  + 1  + 1
tion, the valve dead band produces a steady-state error in the  4   40 
G( s ) = .
system response in the absence of an integrator. For a zero  s  s 
s + 1  + 1
100   300  (27)
Maintaining a low-order controller To satisfy the tracking specification, a pre-
was an important factor in the design filter, F ( s ), is required to place the closed-
loop frequency response in the specified
process. range, betweenTu ( s ) andTl ( s ). Fig 5(a) shows
the frequency response of the closed-loop
steady-state error, L0 must contain an integrator. Thus, a pos- system without a prefilter where the designed controller
sible loop can be obtained by cascading an integrator. Fur- (27) only guarantees that the variation in the system’s mag-
ther, a gain of 0.0065 is required to bring the open-loop nitude will be less than or equal to that allowed. The pre-
transfer function within the specification bounds. Two ze- filter is synthesized by calculating its magnitude bound at
roes should be added to the structure of the controller to sat- each frequency. For example, in Fig. 5(a), consider the fre-
isfy the bounds and to force the nominal loop, L0 , to the quency response of the control system at ω = 50 rad/s. The
right-hand side of the closed boundaries. After several itera- maximum frequency variation, δ, is about 17 dB = 2 − ( −15 ),
tions the controller zeroes were chosen at s = −4 and s = −40. which is less than the amount allowed for this frequency,
At this stage, L0 closely follows the bounds up to ω = 70 rad/s. i.e., δ all = 25 dB =|Tu (i 50 )|−|Tl (i 50 )| = −30 − ( −55 ). The ac-
ceptable magnitude of the prefilter at this frequency is then
found to be −55 − ( −15 ) ≤ |F (i50 )|≤ −30 − (2). By applying
the above procedure for the design frequency range, a suit-
able prefilter is found to be
3 3 4 5

 s 
Analog and Digital u  + 1
Interface Board 150 
6 F (s) = .
 s s 
1  + 1  + 1
2  5.5   9  (28)

The closed-loop Bode plots of the plant with prefilter are


Figure 6. Experimental test station. 1: proportional value, 2: given in Fig. 5(b). Note that all the tracking specifications
pump with pressure regulator, 3: pressure transducer, 4: are met in the frequency domain. The corresponding
incremental encoder, 5: force sensor, 6: environment. closed-loop simulated time responses to a step input of 1000
N are shown in Fig. 5(c). As can be seen, for all cases pertain-
ing to extreme parts of the operating envelope, the specifi-
cations are satisfied.

Experiments

Experimental Setup
The test station consists of a hydraulic unit, a 486/66-based
PC equipped with a Metrabyte M5312 quadrature incre-
mental encoder card, and a DAS-16 analog-to-digital (A/D)
conversion card (Fig. 6). The pump provides constant op-
erational supply pressure up to ≈1000 psi. The hydraulic
valve is a low-cost closed-center four-way proportional
valve. The positioning of the valve spool is based on the
Figure 7. Close-up photograph of the environment. pulse-width modulation principle. A spring is used to rep-

74 IEEE Control Systems Magazine April 2001


resent the environment (Fig. 7); replacing the spring Conclusions
changes the stiffness of the environment. A force trans- This article has described the application of the QFT tech-
ducer with a capacity of 1000 lb (4.5 kN) is inserted be- nique to the development of a force controller for hydraulic
tween the environment and the piston rod. The 12-bit A/D actuators. A linear parametrically uncertain fourth-order
conversion allows a force resolution of 10 N when the full model was developed to represent the relation between the
range is used. Three pressure transmitters read pump, control signal and the force acting on the environment. Plant
supply line, and return line pressures with ±1% accuracy.
An incremental encoder with sensing resolution of 0.06
mm reads the displacement of the cylinder piston. The 600
control signal, generated by the control algorithm, is con-
verted to an analog signal by the A/D card and then trans- 500
Tu
mitted to the hydraulic valve amplifier. The low-cost valve
operates within the range ±1.8 V and has a dead band of 400
Tl

Force (N)
±10% (±0.15 V) within which the actuator does not move.
No attempt was made to eliminate or bypass these 300
nonideal effects due to the intended application.
200
Reference Force = 500 N
Results Ke = 50 kN/m
100
The controller described by transfer function (27) was Ke = 100 kN/m
descretized and implemented on the experimental test
0
stand. The sampling frequency for the controller was 200 Hz. 0 0.5 1 1.5 2 2.5 3 3.5 4
Time (s)
Several experiments were performed to study the effects of (a)
variations of environmental stiffness, supply pressure, and
0.5
force setpoint. First, the variation of environmental stiffness
was studied by using different springs. Two different springs 0.45
Ke = 50 kN/m
with stiffnesses of 50 and 100 kN/m were used for this pur- 0.4
Ke = 100 kN/m
pose. The results are shown in Fig. 8. As can be seen, the 0.35
Control Signal (V)

controller is capable of handling the changes in environ- 0.3


mental stiffness and the steady-state errors are small. With
0.25
reference to Fig. 8(a), responses exhibit initial delays,
0.2
mainly due to the dead band (±10%) and dry friction in the
low-cost valve used in this experiment. Consequently, the 0.15
responses do not precisely fit within the design specifica- 0.1
tion bounds shown with dotted lines. Indeed, the effect of 0.05
valve dead band was not incorporated into the QFT design 0
procedure and was only handled by cascading an integrator 0 0.5 1 1.5 2 2.5 3 3.5 4
Time (s)
in the controller. The actuator displacement and the control (b)
signal for such a trial are shown in Fig. 8(b) and (c), respec-
tively. The control signal is smooth, but it contains high-fre- 0.012
quency oscillations that originated from the noise in the
force sensor. Fig. 9 compares the test results for two differ- 0.01
ent reference forces (500 N and 1000 N) and with similar en-
Displacement (m)

0.008
vironmental stiffness (50 kN/m). As shown in this figure,
despite changing the loading condition, the system’s rise
0.006
time did not change considerably. The ability of the control-
ler to cope with pump pressure variations was also tested.
0.004
Typical results are shown in Fig. 10, where the pump pres-
sure was varied 100%. As can be seen, the control effort is re- Ke = 50 kN/m
0.002
duced for the higher pump pressure. Finally, a test was Ke = 100 kN/m
arranged to investigate the repeatability of the controller
0
and the effect of long-term system operation on the re- 0 0.5 1 1.5 2 2.5 3 3.5 4
sponse. A 1000-N step response was performed once and Time (s)
(c)
then again after the machine had been in continuous oper-
ation for more than one hour. As shown in Fig. 11, there is Figure 8. Step responses with different environmental stiffnesses
no noticeable difference between the two responses. (experiment): (a) force, (b) control signal, and (c) displacement.

April 2001 IEEE Control Systems Magazine 75


uncertainty was quantified within the linear model to com- taining a low-order controller was an important factor in the
pensate for time-varying dynamics and to allow for variations design process to ensure easy implementation and demon-
in the environmental stiffness, hydraulic fluid flow and pres- strate its suitability for industrial applications.
sure gains, actuator compliance, and many other parameters The QFT controller designed here was implemented suc-
specific to the hydraulic actuator components. A robust con- cessfully on an industrial hydraulic actuator equipped with
troller was then designed that, along with a prefilter, main- a low-cost proportional valve. The transparency of the de-
tains a satisfactory force control performance against the sign process facilitated the quick adjustment of the control-
environment despite a wide range of uncertainty. Main- ler for experimental implementation. Several tests were
performed to show the force-regulating ability of the devel-
oped controller. In particular, the experimental results dem-
1200
onstrated the robustness of the QFT controller for up to
1000 100% variations in environmental stiffness, supply pres-
sure, and reference force. Good performance was also ob-
800 tained despite significant variations in actuator dynamics.
Force (N)

The single-loop fixed-gain controller was easy to implement,


600
required very little computational effort, needed no online
tuning or gain scheduling, and resulted in good perfor-
400
mance in both transient and steady-state periods. Knowl-
200 edge of the lower and upper bounds of uncertain parameters
Ke = 50 kN/m
was the only requirement for the controller design.
0 In summary, the results of this work, which represent a
0 0.5 1 1.5 2 2.5 3 3.5 4
Time (s) contribution to the published work on the application of
QFT, provide insight into the potential and effectiveness of
Figure 9. Step responses for different reference forces the technique for the design of robust force controllers for
(experiment). hydraulic actuators. In particular, we have shown that it is
feasible to employ a single fixed-gain force controller for a
1200

1000 1200

800 Stiffness = 75 kN/m 1000


Force (N)

600 800
Force (N)

400 Pump Pressure = 500 psi 600 Stiffness = 75 kN/m


Pump Pressure = 1000 psi Pump Pressure = 750 psi
200 400

200
0 0.5 1 1.5 2 2.5 3 3.5 4
Time (s) 0
(a) 0 0.5 1 1.5 2 2.5 3 3.5 4
Time (s)
0.6 (a)
0.6
0.5 Pump Pressure = 500 psi
Pump Pressure = 1000 psi 0.5
Control Signal (v)

0.4
Control Signal (v)

0.4
0.3
0.3
0.2
0.2
0.1
0.1
0 0.5 1 1.5 2 2.5 3 3.5 4
Time (s) 0
0 0.5 1 1.5 2 2.5 3 3.5 4
(b) Time (s)
(b)
Figure 10. Step force responses for different supply pressures
(experiment). Figure 11. Repeatability test (experiment).

76 IEEE Control Systems Magazine April 2001


hydraulic actuator, and that control performance can be in- [18] S.N. Sheldon and S.J. Rasmussen, “Development and first successful
flight test of a QFT flight control system,” in Proc. IEEE National Aerospace and
sensitive to structured plant variation. It also became ap- Electronics Conf., NAECON, Dayton, OH, 1994, pp. 629-636.
parent that the uncertain linear fourth-order model
[19] M.B. Leahy, D.E. Bossert, and P.V. Whalen, “Robust model-based control:
adequately characterizes the major features of the class of An experimental case study,” in Proc. IEEE Conf. on Robotics and Automation,
hydraulic systems under investigation. Sacramento, CA, 1990, pp. 1982-1987.
In future work, we plan to use the QFT method to design [20] S.J. Rober, Y.C. Shin, and O.D.I. Nwokah, “A digital robust controller for
automatic force controllers for heavy-duty hydraulic equip- cutting force control in the end milling process,” ASME J. Dynam. Syst., Mea-
ment such as excavators. These machines strongly interact sure., Contr., vol. 119, no. 2, pp. 146-152, 1997.
with environments, and accurate models for the interac- [21] M. Pachter, C.H. Houpis, and K. Kang, “Modeling and control of an
tions are difficult to obtain. Currently, we are investigating electrohydrostatic actuator,” Int. J. Robust Nonlinear Contr., vol. 7, no. 6, pp.
591-608, 1997.
the feasibility of assuming the environmental force as a
combination of disturbance and unstructured uncertainty [22] D.F. Thompson and G.G. Kremer, “Quantitative feedback design for a
variable-displacement hydraulic vane pump,” in Proc. ACC, Albuquerque,
to be handled by the QFT controller.
NM, 1997, pp. 1061-1064.

[23] R. Volpe and P. Khosla, “Theoretical analysis and experimental verifica-


References tion of a manipulators/sensor/environment model for force control,” in Proc.
[1] A. Alleyne, “Nonlinear force control of an electrohydraulic actuator,” in IEEE Conf. Systems, Man, Cybernetics, Los Angles, CA, 1990, pp. 784-790.
Proc. Japan/USA Symp. Flexible Automation, Boston, MA, 1996, pp. 193-200.
[24] S.D. Eppinger and W.P Seering, “On dynamic models of robot force con-
[2] F. Conrad and C.J.D. Jensen, “Design of hydraulic force control systems trol,” Proc. IEEE Conf. Robotics and Automation, San Francisco, CA, 1986, pp.
with state estimate feedback,” in Proc. IFAC 10th Triennial Congr., Munich, 29-34.
Germany, 1987, pp. 307-312.
[25] H.E. Merritt, Hydraulic Control Systems. New York: Wiley, 1967.
[3] N. Niksefat and N. Sepehri, “Robust force controller design for an
electrohydraulic actuator based on nonlinear model,” in Proc. IEEE Conf. Ro- [26] K. Ziaei, “Development of a nonlinear adaptive control scheme for hy-
botics and Automation, Detroit, MI, 1999, pp. 200-206. draulic actuator,” M.Sc. thesis, Dept. Mech. Eng., Univ. of Manitoba, Winni-
peg, Canada, 1998.
[4] R.F. Pannett, P.K. Chawdhry, and C.R. Burrows, “Alternative robust con-
trol strategies for disturbance rejection in fluid power systems,” in Proc. ACC, [27] H. Seraji and R. Colbaugh, “Force tracking in impedance control,” Int. J.
San Diego, CA, 1999, pp. 739-743. Robot. Res., vol. 16, no. 1, pp. 97-117, 1997.
[5] J. Watton, “On linearized coefficients for an underlapped servo-valve cou- [28] S. Jayasuriya, “Frequency domain design for robust performance under
pled to a single-rod cylinder,” ASME J. Dynam. Syst., Measure., Contr., vol. 112, parametric, unstructured, or mixed uncertainties,” ASME J. Dynam. Syst.,
no. 4, pp. 794-796, 1990. Measure., Contr., vol. 115, no. 2, pp. 439-451, 1993.
[6] J. Yu, Z. Chen, and Y. Lu, “The variation of oil effective bulk modulus with [29] I.M. Horowitz, “Survey of quantitative feedback theory (QFT),” Int. J.
pressure in hydraulic systems,” ASME J. Dynam. Syst., Measure., Contr., vol. Contr., vol. 53, no. 2, pp. 255-291, 1991.
116, no. 1, pp. 146-150, 1994.
[7] Y.N. Chen, C.B. Lee, and C.H. Tseng, “A variable-structure controller de-
sign for an electrohydraulic force control servo system,” J. Chinese Soc. Mech.
Navid Niksefat received the B.Sc. degree from Sharif Uni-
Eng., vol. 11, no. 6, pp. 520-526, 1990. versity of Technology, Iran, and the M.Sc. degree from
[8] H. Sun and G.T.C. Chiu, “Nonlinear observer based force control of Isfahan University of Technology, Iran, in 1994 and 1997, re-
electrohydraulic actuators,” in Proc. ACC, San Diego, CA, 1999, pp. 764-768. spectively, both in mechanical engineering. Since January
[9] R. Liu and A. Alleyne, “Nonlinear force/pressure tracking of an 1998, he has been with the Experimental Robotics and
electrohydraulic actuator,” in Proc. IFAC 14th Triennial Congr., Beijing, China, Teleoperation Laboratory, University of Manitoba, Canada,
1999, pp. 469-474.
where he has held teaching and research assistantship posi-
[10] G. Wu, N. Sepehri, and K. Ziaei, “Design of a hydraulic force control sys-
tions. His current research interests include robust control
tem using a generalized predictive control algorithm,” Proc. Inst. Elec.
Eng.-Contr. Theory Applicat., vol. 145, no. 5, pp. 428-436, 1998. and control of discontinuous and nonlinear systems, with
[11] L. Laval, N.K. M’Sirdi, and J.C Cadiou, “H ∞ force control of a hydraulic applications to robotics and hydraulic systems. He re-
servo-actuator with environmental uncertainties,” in Proc. IEEE Conf. Robot- ceived several awards and a fellowship from the University
ics and Automation, Minneapolis, MN, 1996, pp. 1566-1571. of Manitoba during 1999-2001.
[12] G. Ferretti, G. Magnani, and P. Rocco, “On the stability of integral force
control in case of contact with stiff surface,” ASME J. Dynam. Syst., Measure.,
Contr., vol. 117, no. 4, pp. 547-553, 1995. Nariman Sepehri joined the Department of Mechanical and
[13] I.M. Horowitz, Quantitative Feedback Design Theory (QFT). Boulder, CO:
Industrial Engineering at the University of Manitoba in 1991,
QFT Publications, 1992. where he is currently a Professor. He received an M.Sc. in
[14] J.J. D’Azzo and C.H. Houpis, Linear Control System Analysis and Design. computer-aided process planning in 1986 and a Ph.D. in ro-
New York: McGraw-Hill, 1988. botics and controls in 1990, both from the University of Brit-
[15] O. Yaniv, Quantitative Feedback Design of Linear and Nonlinear Control ish Columbia. His areas of interest are mathematical
Systems. Norwell, MA: Kluwer, 1999. modeling and real-time simulation, robust control design,
[16] S. Jayasuriya, O. Yaniv, O.D.I. Nwokah, and Y. Chait, “Benchmark prob- and fluid power. An area of research interest is the automa-
lem solution by quantitative feedback theory,” J. Guidance, Contr., Dynam.,
tion of heavy-duty industrial (forest and mining) machinery
vol. 15, no. 5, pp. 1087-1093, 1992.
using robotics and advanced control technology. He is a
[17] S.G. Breslin and M.J. Grimble, “Longitudinal control of an advanced com-
bat aircraft using quantitative feedback theory,” in Proc. ACC, Albuquerque, member of the American Society of Mechanical Engineers
NM,1997, pp. 113-117. and the IEEE.

April 2001 IEEE Control Systems Magazine 77

You might also like