You are on page 1of 10

Tip position control of a lightweight exible manipulator using a fractional order controller

C.A. Monje, F. Ramos, V. Feliu and B.M. Vinagre Abstract: A new method to control single-link lightweight exible manipulators in the presence of payload changes is proposed. Undoubtedly, the control of this kind of structures is nowadays one of the most challenging and attractive research areas, being remarkable its application to the aerospace industry, among others. One of the interesting features of the design method presented here is that the overshoot of the controlled system is independent of the tip mass. This allows a constant safety zone to be delimited for any given placement task of the arm, independent of the load being carried, thereby making it easier to plan collision avoidance. Other considerations about noise and motor saturation issues are also presented. To satisfy this performance, the overall control scheme proposed consists of three nested control loops. Once the friction and other nonlinear effects have been compensated, the inner loop is designed to give a fast motor response. The middle loop simplies the dynamics of the system and reduces its transfer function to a double integrator. A fractional derivative controller is used to shape the outer loop into the form of a fractional order integrator. The result is a constant phase system with, in the time domain, step responses exhibiting constant overshoot, independent of variations in the load, and robust, in a stability sense, to spillover effects. Experimental results are shown, when controlling the exible manipulator with this fractional order derivator, that prove the good performance of the system.

Introduction

During the last two decades, a considerable interest has been attracted to the control of lightweight exible manipulators, which is becoming one of the most challenging research areas of robotic control. The necessity of this kind of robots arises from new robotic applications that require lighter robots that can be driven using smaller amounts of energy, such as the aerospace industry, where weight has to be minimised, or mobile robotics, where power limitations imposed by battery autonomy have to be taken into account. In addition, collisions of this type of robots present remarkably less destructive effects than those caused by traditional robots, as the kinetic energy of the movement is transformed into potential energy of deformation at the moment of impact. This fact allows us to perform some control strategy over the actuator before any damage takes place or, at least, to minimise it, which may lead, in a not very far future, to a robot human cooperation without the actual dangers in case of malfunction, which is an emergent topic of research interest [1]. The challenge of controlling vibrations in exible structures, such as robotic manipulators, has been approached with very different methods, from classical schemes [2], to nonlinear methods such as sliding control [3] or neural networks [4]. A recent published survey [5] gives a complete, detailed overview of all the work developed in this
# The Institution of Engineering and Technology 2007 doi:10.1049/iet-cta:20060477 Paper rst received 8th November 2006 and in revised form 14th February 2007 C.A. Monje and B.M. Vinagre are with the Industrial Engineering School, University of Extremadura, Badajoz, Spain F. Ramos and V. Feliu are with the Superior Technical School of Industrial Engineers, University of Castilla-La Mancha, Ciudad Real, Spain E-mail: cmonje@ing.uc3m.es IET Control Theory Appl., 2007, 1, (5), pp. 1451 1460

eld since the late 1970s. However, in spite of all the research devoted to modelling and controlling these kind of robots, there is no universal solution for the control, which is clearly demonstrated by the number of recent papers presenting new improved solutions for vibration control. The present work shows a new and simple control scheme for single-link exible arms with a variable payload, based on the use of a fractional order derivative controller. This scheme uses measurement of the link deection provided by a strain gauge placed at the base of the link. This sensorial system lets us construct control schemes that are more robust than those based on accelerometer measurements (as was demonstrated in [6], too), being strain gauges simpler to instrument. The general control scheme proposed in this paper consists of three nested loops (see Fig. 1): 1. An inner loop that controls the position of the motor. This loop uses a classical proportional derivative (PD) controller to give a closed loop transfer function close to unity. 2. A simplifying loop using positive unity-gain feedback. The purpose of this loop is to reduce the dynamics of the system to that of a double integrator. 3. An outer loop that uses a fractional order derivative controller to shape the loop and to give an overshoot independent of payload changes. In Fig. 1, um is the motor angle, ut the tip-position angle, Gb(s) the transfer function of the beam and Ri(s) and Re(s) the inner and outer loop controllers, respectively. The design of the rst two loops follows [7]. The fractional order control strategy of the outer loop, which is based on the operators of fractional calculus (see [8]), is presented in this paper. A former use of this strategy was discussed in [9] considering a more ideal case and only simulation
1451

controls the motor, J the motor inertia, n the viscous friction ^ coefcient, G Coul the coupling torque between the motor ^ and the link and G Coul the Coulomb friction. From now on, we will suppose that the Coulomb friction is negligible or is compensated by a term (see [13]) of the form  G _ Coul ^ m) sign(u (3) K ^ as shown in Fig. 2a, where G Coul is an estimation of the Coulomb friction value. On the other hand, the coupling torque equation between the motor and the link is VCoul

Fig. 1 Proposed general control scheme

results. The state of the art in fractional order control can be found in [10 12]. For a better understanding of this work, it is organised as follows. First, the physical model of the single-link exible manipulator is presented in Section 2, followed by a description of the general control scheme and the three nested loops in Section. 3 The design of the outer loop is explained in detail in Section 4 and the effect of higherorder dynamics is discussed in Section 5. Next, Section 6 presents the results obtained from the test of the control strategy in an experimental platform. Finally, some relevant concluding remarks are drawn in Section 7. 2 Modelling of the single-link exible manipulator The single-link lumped mass model that will be used in this paper is well known [6]. Particularly, a single tip mass that can rotate freely (no torque is produced at the tip) will be adopted for the description of the link dynamics [6]. The effect of the gravity is assumed negligible as the arm moves in a horizontal plane. The motor has a reduction gear with a reduction relation n. The magnitudes seen from the motor side of the gear will be written with an upper hat, whereas the magnitudes seen from the link side will be denoted by standard letters. The dynamics of the link is described by c(um ut ) ml u t
2

Gcoup c(um ut )

(4)

^ G =n nu and G and nally, the conversion equations u complete the dynamic model. Laplace transform is applied to (1) leading to the following transfer function G b ( s)

u t ( s) v2 2 0 2 um (s) s v0

(5)

2 where v0 is the natural frequency of the link v2 0 c/ml , which is mass-dependent. Combining all the previous equations, the transfer functions of the robot are

um (s) Kn(s2 v2 0) (6) 2 2 3 2 2 2 V ( s) s[Jn s nn s (Jn v0 c)s nn2 v2 0] u t ( s) Knv2 0 (7) 2 3 2 2 2 2 V (s) s[Jn s nn s (Jn2 v2 0 c)s nn v0 ]
It is evident that these robot equations are massdependent (so is v0) and therefore changes in the mass will affect the system behaviour. 3 General control scheme

(1)

where m is the mass at the end, l and c are the length and the stiffness of the bar, respectively, um is the angle of the motor and ut is the angle of the tip. The dynamics of the motor with a closed-loop current control system (where the voltage V is proportional to the current output) is given by _ ^ ^ ^ m nu ^m G KV J u coup GCoul (2)

where K is the motor constant, V the voltage signal that

The general control scheme is shown in Fig. 1, where it can be seen that it is composed of three nested loops: inner loop, simplifying loop and outer loop, as commented previously. The features of the inner and outer loops have been previously detailed in [6], whereas the simplifying loop is now included to cope with the fractional order control strategy. Basically, this scheme allows us to design the loops separately, making the control problem simpler and minimising the effects of the inaccuracies in the estimation of Coulomb and viscous frictions in control performance (as shown in [13]).

Fig. 2 Block diagram of the loops


a Block diagram for the inner loop b Block diagram for the simplifying loop
1452
IET Control Theory Appl., Vol. 1, No. 5, September 2007

3.1

Inner loop

It is important to remark that the coupling torque is compensated within the inner loop by a term of the form Vcoup 1 G Kn coup (9)

The inner control loop, shown in Fig. 2a, fastens the dynamic behaviour of the motor. This purpose is achieved by means of a standard PD controller with proportional constant Kp and derivative constant Kv , which is tuned to make the motor dynamics critically damped. The second-order critically damped expression obtained for the motor loop is

M ( s)

u m ( s) 1 ur ( s ) (1 g s) 2 m

(8)

Previous experimental works have proven the correctness of this assumption in direct driven motors, and motors with reduction gears as well [7]. It has also been demonstrated that, in the case of motors with gears, the effect of the coupling torque is very small compared with the motor inertia and friction, as its value is divided by n [6]. 3.2 Simplifying loop

where ur m(s) is the reference angle for the motor and g the p motor dynamics constant, which is given by g (J =Kp K ). Theoretically, it is possible to make the motor dynamics as fast as desired by simply making g ! 0, but a very demanding speed would saturate the motor, with the subsequent malfunction of the controlled system. This fact implies that, although the motor dynamics can be made quite fast, we cannot consider it negligible in general. Effects of this dynamics can be studied by dividing the overall system into two time-scale subsystems: the fast dynamics subsystem dened by the inner control loop and the slow dynamics subsystem dened by the exible link. Then, interactions between both subsystems are studied. Singular perturbation techniques [14] dene the general framework to carry on this study. Application of these techniques to exible arms can be found in [15], where the arm dynamics is divided into a slow subsystem (which includes motor dynamics) and a fast subsystem (which includes high-order vibration modes). A recent work based on this approach is presented in [16], where the dynamics of the highest vibration modes are neglected in the controller design but the maximum possible gain in the L2 sense of these removed dynamics is taken into account in such a design. It is possible to modify these techniques in order to cope with the exible arm control problem dened in this paper. In this sense, the effects of the servocontrolled motor fast dynamics on the slow dynamics dened by the exible link and the outer controller can be quantied and used to adequately tune the outer controller. But the main purpose of this paper is to show the feasibility of a robust fractional controller designed by using a very simple technique that does not take into account this fast dynamics. We will show in Section 4.5 that the effects of the inner loop fast dynamics can be reduced by adding a proportional term (tuned easily by simulations) to the fractional controller designed with our robust control technique. Then, a more elaborate analysis is not needed to achieve acceptable results with our arm. Tables 1 and 2 show the parameters of the motor-gear set and the exible link used for the experimental tests, respectively.
Table 1: Data of the motor-gear set
g
0.022 Kp 1 Kv 0.025 J, (kg m 2) 24.24 . 10
24

As commented previously, the response of the inner loop (position control of the motor) is signicantly faster than the response of the outer loop (position control of the tip). The motor position is rst supposed to track the reference position with negligible error and the motor dynamics will be considered later. That is, the dynamics of the inner loop can be approximated by 1 when designing the outer loop controller. Taking this into account, a strategy for simplifying the dynamics of the arm, shown in Fig. 2b, is proposed. For the case of a beam with only one vibrational mode, a simplifying loop can be implemented that reduces the dynamics of the system to a double integrator by simply closing a positive unity-gain feedback loop around the tip position (b 1). Then, the equation relating the output and input of the loop is

u t ( s)

v2 1 0 u(s) 2 P(s) s2 s

(10)

where P(s) represents disturbances with the form of a rstorder polynomial in s, which models initial deviations in tip position and tip velocity [6]. In (10), the dynamics of the arm has been reduced to a double integrator dynamics, simplifying the control strategy proposed in this work, as will be seen later. The stability study by using Nyquist diagrams shows that the condition b 1 is not critical to obtain stable control systems, being sufcient to implement a feedback gain close to 1. 3.3 Outer loop

The block diagram for the outer loop used in this work is shown in Fig. 3. As it is observed in the scheme, an estimation of the tip position, ue t , is used to close the loop. We actually feed back the deformation measurements of the two strain gauges, placed at the base of the link in a halfbridge, 2-active-gauges conguration, to control the arm. These sensors provide the value of the coupling torque

n, (N m s) 51.66 . 10
24

K, (N m/V ) 3.399

n 50

Table 2: Data of the exible link


c, (N m) 443.597 m, (kg) 1.9 l, (m) 0.866 r, (m) 0.008 E, (GPa) 68.9 I, (m 4) 1.86e-9 Ks 2.11 1453

IET Control Theory Appl., Vol. 1, No. 5, September 2007

frequency specications found for the system, as will be shown later. 4 Design of the outer loop controller Re(s)

With the inner and simplifying loops closed, the reduced diagram of Fig. 4 is obtained, which is based on (10). From this diagram, the expression for the tip position is
Fig. 3 Basic scheme of the outer control loop

ut (s)

1 1 s2 =R
2 e ( s) v 0

ur t ( s)

1 1 s2 = R
2 e (s)v0

P ( s) Re (s)v2 0 (13)

Gcoup between the arm and the motor by means of expression Gcoup
1 2:00 EI e 2 Ks r (11)

where Ks is the gauge factor provided by the manufacturer, r the outer beam radius, E is Youngs modulus of the beam material and I the cross-section inertia of the beam. These values can be found in Table 2. On the other hand, e represents the deformation measured by the strain gauges, which is proportional to a voltage signal and calibrated with a dynamic strain amplier. The 1=2 gain is due to the chosen conguration, which doubles the sensed value while remains insensitive to temperature variations and cancels the compressive/tensile strain. Coupling torque is used to decouple motor and link dynamics and estimate tip position [6]. Combining (1), (4) and the fundamental frequency denition, we can obtain the relation between Gcoup and the output um , which yields G G ( s) cs2 s2 v 2 0 (12)

The controller Re(s) has a 2-fold purpose. One objective is to obtain a constant phase margin in the frequency response; in other words, a constant overshoot in time response to a step reference for varying payloads. The other is to remove the effects of the disturbance, represented by the initial conditions polynomial, on the steady state. To attain these objectives, most authors propose the use of some kind of adaptive control scheme (see [7]). We propose here a fractional order derivative controller with enhanced robustness properties to achieve the above two objectives, without needing any kind of adaptive algorithm. Some methods have been developed in the last years to tune fractional propotional integrative derivative (PID) controllers with robustness properties to changes in one parameter (typically the plant gain) [17, 18]. These methods are based on optimisation procedures. Our approach is much simpler as it is specically tailored to our particular arm dynamics, leading to very straightforward tuning rules. 4.1 Condition for constant phase margin

Given the values of um and Gcoup , experimentally measured, the value of ut can be estimated according to (4) by ue t um 2 Gcoup/c. This estimated tip angle is used to close the control loop, as shown in Fig. 3, where H(s) 2 1/c. The main purpose of this work is to design the outer controller Re(s) (see Fig. 3) so that the time response of the controlled system has an overshoot independent of the tip mass and the effects of disturbances are removed. This will lead to the use of a fractional derivative controller, as will be detailed in the next section. Besides, in this particular case, the outer controller will be designed in the frequency domain for the specications of phase margin (damping of the response) and crossover frequency (speed of the response). In order to guarantee a critically damped response (overshoot Mp 0), a phase margin wm 76.58 is selected. Besides, the response is desired to have a rise time around 0.3s, so the crossover frequency is xed to vcg 6 rad/s. The crossover frequency denes the speed of response of the closed-loop system. The practical constraint that limits the speed of response of the arm, and hence the value of vcg , is the maximum torque provided by the motor. This maximum torque limits the speed of response of the inner motor loop and hence its bandwidth. The control scheme proposed here works ideally if the dynamics of the inner loop is negligible. Consequently, its bandwidth must be much larger than the desired crossover frequency vcg . However, it must be taken into account that the experimental results to be presented in this work show the behaviour of the arm assuming non-negligible inner loop dynamics, since the value of the torque provided by the motor limits the speed of response of this loop. This fact may change slightly the nal
1454

The condition for a constant phase margin can be expressed as   v2 0 arg Re (jv) constant 8v (14) (jv)2 and the resulting phase margin wm is

wm arg [Re ( jv)]

(15)

For a constant phase margin 0 , wm , p/2, the controller that achieves this must be of the form Re (s) Kc sa , a 2 w p m (16)

so that 0 , a , 1. This Re(s) is a fractional derivative controller of order a. The two denitions used for the general fractional integro-differential operation are the Gru nwald Letnikov (GL) denition and the Riemann Liouville (RL) denition [8]. The GL denition is
a a D t f (t )

lim h
h !0

[h] X j0

t a

  a f (t jh) ( 1) j
j

(17)

Fig. 4 Reduced diagram for the outer loop


IET Control Theory Appl., Vol. 1, No. 5, September 2007

where [.] means the integer part, whereas the RL denition is 1 dn t f ( t) a dt (18) a D t f (t ) n G(n a) dt a (t t)an1 for (n 2 1 , a , n) and where G(.) is the Eulers gamma function. For convenience, Laplace domain notion is usually used to describe the fractional integro-differential operation. The Laplace transform of the RL fractional derivative/integral (18) under zero initial conditions for order a (0 , a , 1) is given by
a +a L{a D+ F ( s) t f (t)} s

or, equivalently, a phase margin between 908 and 08. The second one is the gain Kc to adjust the crossover frequency, or, equivalently, the speed of the response for a nominal payload. Note that increasing a decreases the overshoot but increases the time required to correct the disturbance effects (see [20]). 4.4 Controller design

As commented above, the design of the controller thus involves the selection of two parameters: a, the order of the derivative, which determines: (a) the overshoot of the step response, (b) the phase margin or (c) the damping. Kc , the controller gain, which determines for a given a: (a) the speed of the step response or (b) the crossover frequency. These parameters can be selected by working in the complex plane, the frequency domain or the time domain. In the frequency domain, the selection of the parameters of the fractional order derivative controller can be regarded as choosing a xed phase margin by selecting a, and choosing a crossover frequency vcg , by selecting Kc for a given a. That is

(19)

Taking this notation into account, Re(s) corresponds to a fractional derivative controller of order a. 4.2 Condition for removing the effects of disturbance From the nal value theorem, the condition to remove the effects of the disturbance is
s !0 1

lim

1 P(s) 0 2 (s2 =Re (s)v2 R ) e ( s) v 0 0

(20)

Substituting Re(s) Kcs a and P(s) as b (initial tip position and velocity errors different from zero), this condition becomes b a lim s 0 s!0 1 (s2a =Kc v2 ) Kc v2 0 0
s!0 1

2 w , p m

Kc

vcg v2 0

(26)

(21) (22)

lim

1 a 1a s 0 2 ) (s2a =Kc v2 K c v0 0

which implies that a , 1. 4.3 Ideal response to a step command

Assuming that the dynamics of the inner loop can be approximated by unity and that disturbances are absent, the closed-loop transfer function with controller (16) is Fcl (s)

ut 1 Kc v2 0 ur s2a Kc v2 1 (s2 =Re (s)v2 t 0) 0

(23)

which exhibits the form of Bodes ideal loop transfer function [19]. The corresponding step response is   Kc v2 1 0 u t (t ) L s(s2a Kc v2 0)
2 a 2a E2a,3a ( Kc v2 ) Kc v2 0t 0t

According to Table 2, where the parameters of the exible manipulator are presented, the fundamental frequency of the system is v0 17.7 rad/s. The frequency specications required for the controlled system, commented previously, are: phase margin wm 76.58, and crossover frequency around vcg 6 rad/s. Therefore the parameters of the fractional derivative controller are a 0.85 and Kc 0.02. With this controller, and under the assumption of negligible inner loop dynamics, the Bode plots obtained for the open-loop system are shown in Fig. 5a, where it can be observed that at the crossover frequency vcg 6 rad/s the phase margin is wm 76.58, fullling the design specications. The simulated step responses of the controlled system for m 0.6, 1.9, 3.2 and 6 kg are shown in Fig. 5b. It is observed that the overshoot of the response remains constant to payload changes, being Mp 0, fullling the robustness purpose. For the nominal mass (m 1.9 kg), a rise time tr 0.3 s is obtained. This controller has been implemented as described in Section 4.6, except for the constant k 0.25, which has been introduced later to compensate the effects of the nonnegligible inner loop dynamics, as explained next. 4.5 Effect of the non-negligible inner loop dynamics In the practical case presented in this work, the dynamics of the inner loop is not negligible, being given by the transfer function M ( s)

(24)

where Ed,d1(2 At d) is the two-parameter Mittag Lefer function [8]. The overshoot is xed by 2 2 a, which is independent of the payload, and the speed by Kcv2 0, that is, by the payload and the controller gain. In fact, notice that this expression can be normalised with respect to time by
2a 2a ut (tn ) tn E2a,3a ( tn )

(25)

1/(22a) . This equation shows that the where tn t(Kcv2 0) effect of a change in the payload implies a change in v0 that only means a time scaling of the response ut(t ). To obtain a required step response, it is then necessary to select the values of two parameters. The rst one is the order a to adjust the overshoot between 0 (a 1) and 1 (a 0),

um (s) 1 ur ( s ) (1 gs)2 m

(27)

with g 0.022. Notice that M(s) is independent of the value of the tip payload as its effects on the motor dynamics are removed by the compensation term Vcoup in (9) based on the measurement of the motor-beam coupling torque. The introduction of M(s) implies that the response of the
1455

IET Control Theory Appl., Vol. 1, No. 5, September 2007

Fig. 5 Bode plots and simulated responses with fractional order derivative
a Bode plots of the open-loop system with the fractional order derivative, considering negligible inner loop dynamics b Simulated time responses of the system with the fractional order derivator to a step input for different payloads, considering negligible inner loop dynamics

controlled system will be affected by this dynamics, as the simplifying loop does not result in a double integrator anymore. Besides, it is important to remark that step inputs are not very appropriate for robotic systems, being more suitable the use of smoother references to avoid surpassing the physical limitations of the robot, such as the maximum torque allowed to the links before reaching the elastic limit or the maximum feasible control signal value (V for a DC motor-amplier set). For that reason, a fourth order polynomial reference ur t has been used in our case. It has been observed that with the introduction of M(s), the settling time of the response gets longer. To reduce it, a proportional part k is introduced in the controller to make the output converge faster to its reference. However, it must be taken into account that the introduction of this constant affects the frequency response of the system, changing the specications. Therefore there must be a trade-off between the fulllment of the frequency specications and the settling time required, resulting k 0.25 in our case. Then, the nal controller is Re (s) 0:25 0:02s0:85 (28)

discretisation method is used. That is, rst a nitedimensional continuous approximation is obtained and secondly the resulting s-transfer function is discretised. It must be taken into account that the fractional derivative s 0.85 has been implemented as s 0.85 s . s 20.15 (s 1/s 0.15), that is, an integer derivative plus a fractional integrator. This way, the resulting open-loop system in the ideal case would 2 20.15 2 be Re(s)(v2 , guaranteeing the cancella0/s ) (v0/s)s tion of the steady-state position error due to the effect of the pure (integer) integral part. Therefore only the fractional part Rd(s) s 20.15 has been approximated. To obtain a nite-dimensional continuous approximation of the fractional integrator, a frequency domain identication technique is used, provided by the Matlab function invfreqs. An integer order transfer function that ts the frequency response of the fractional order integrator Rd in the range v [ (1022, 102) is obtained. Later, the discretisation of this continuous approximation is made by using the Tustin rule with prewarp frequency vcg and sample period Ts 0.002 s, obtaining a fth-order digital innite impulse response (IIR) lter 0:1124z5 0:7740z4 2:0182z3 2:5363z2 1:5523z1 0:3725 Rd (z) 0:4332z5 2:6488z4 6:3441z3 7:4747z2 4:3462z1 1

Only a slight modication of the frequency specications is obtained with this controller, resulting vcg 6.6 rad/s and wm 708. The time responses of the system for payload changes are shown in Fig. 6. Masses greater than 3.2 kg have not been considered as they could cause the beam to reach its elastic limit and, hence, they will be neither simulated nor experimented. For the nominal mass, an overshoot Mp 0% is obtained. As far as the robustness is concerned, a slight change in the overshoot of the response appears when the payload changes, due to the effect of the non-negligible inner loop dynamics. However, only a 0.59% variation in the overshoot is obtained for the different masses. 4.6 Fractional order controller implementation

(29)

No physical devices are available to perform the fractional derivatives, and then approximations are needed to implement fractional controllers. These approximate implementations of fractional order controller (FOC) can be classied into either continuous [21 23] or discrete methods [22, 24]. In this particular case, an indirect
1456

Fig. 6 Time responses of the system with controller Re(s), considering non-negligible inner loop dynamics
IET Control Theory Appl., Vol. 1, No. 5, September 2007

Therefore the resulting total fractional order controller is a sixth-order digital IIR lter given by   1 z1 Re (z) 0:25 0:02 Rd (z) (30) Ts 5 Robustness to higher vibration modes

This section studies the robustness of the developed fractional controller to non-modelled higher vibration modes. These modes can inuence the closed-loop system in two ways: (1) they are fed back to the controller and, if they were not taken into account in the controller design, the global system can become unstable and (2) the estimator of the tip position based on expression (4) no longer remains correct, exhibiting high-frequency estimation errors that are fed back to the closed-loop system. In order to avoid these destabilising effects, we propose a lemma based on the next dynamic property. Let us consider the transfer function GG(s) between the motor angle and the motor-beam coupling torque (the other measured variable). Then we say that this transfer function exhibits the interlacing property of the poles and zeros on the imaginary axis if it veries that
2 2 s ( s 42 1 ) ( s 4i ) 2 2 ( s 4n ) 2 2 2 2 (s v0 )(s2 v2 1 ) (s v i ) 2 2 (s vn ) 2 2

Fig. 7 Interlacing property between poles (dashed lines) and zeroes (solid lines) of transfer function GG(s) is numerically demonstrated for a wide range of beam masses

G G ( s)

(31)

where vi21 , 4i , vi , 1 i n, and c . 0. This property is veried by uniform single-link exible manipulators with distributed mass and a payload at the tip as illustrated next. The governing equation of a exible link (Euler Bernoulli equation) can be normalised by dening p T (rl4 =EI ), where r is the mass per unit length and introducing the dimensionless time tn t/T. Consequently, the tip payload is also normalised with respect to the beam mass: mn rl/m. Transfer functions GG(s) are obtained for the normalised beam for different mn ratios and the poles and zeros associated to the rst six modes are calculated. We assume that modelling six modes is enough to study the spillover effects in most exible manipulators. Fig. 7 shows the values of these poles

and zeros for mass ratios ranging from negligible link mass (mn 0.01) to the case of a link mass 10 times larger than the tip payload (mn 10). This gure shows that the aforementioned interlacing property is veried by any uniform beam at least in the specied range of variation of the link mass. The following lemma proves that if this property is veried, the controller proposed in the previous section is robust to non-modelled higher vibration modes (spillover). Lemma 1: Assume that our exible arm veries the interlacing property (31) and that the inner loop dynamics is negligible (M(s) 1). Then any outer loop controller of the form R e ( s) k K c sa , Kc . 0, 0

(32)

combined with a tip position estimator of the form given by

ue t um

Gcoup ^ c

(33)

Fig. 8 Nyquist plots used in the study of spillover


a GG(s) proves to be marginally stable when interlacing property is fullled b Addition of the controller Re(s) and the simplifying loop subtracts phase, achieving a stable behaviour for controlled system
IET Control Theory Appl., Vol. 1, No. 5, September 2007

1457

^ veries where parameter c ^!c c keeps stable the closed loop system. Proof: Fig. 3 shows the block diagram of the outer control loop. Operating this block diagram, we obtain the equivalent transfer function (34)

that the Nyquist plot does not embrace the point (2 1,0), it must be veried that
v!1 c ^

lim

1 GG (s)( 1 R e (s)) ! 1

(36)

Gcoup (s) GG (s) 1 ^ Qr ( s ) ( 1 1 = c G G s)( 1 Re (s)) t

(35)

Assuming that Re(s) is of the form (32) and taking into account (31), it easily follows that condition (36) becomes ^ ! c, and expression (34) is proven. c In addition, if Re(s) is of the form (32), after some operations we have that 1 1 j(jv) ( 1 R e ( jv)) ^ c k (1 k ) Kc va 2 2 (1 2k ) cos (p=2)a Kc v a jKc v sinp=2a 2 2a ^ (k 2 2Kc k va cos (p=2)a Kc v ) c

If GG(s) veries the above interlacing property, the alternation between poles and zeros of expression (31) produces a Nyquist diagram of GG( jv) of the form shown in Fig. 8a. It exhibits as many half-turns in the innity as vibration modes has the transfer function [as many as terms (s 2 v 2 i ) are in the denominator of this transfer function]. This plot shows that the closed-loop system associated to 1 ^ ( 1 R GG(s) is marginally stable. The product 1=c e (s)) subtracts phase to the system from zero, when frequency is very small, to 1808, when frequency tends to innity, hence progressively rotating the Nyquist of GG( jv), but never crossing the negative x-axis, as Fig. 8b shows for a three vibrational modes example. In order to guarantee

(37)

The imaginary component of this expression is negative 8v ! 0 provided that 0 a 1 and K ! 0. Then /j( jv) 0 8v ! 0, and it subtracts phase from GG( jv) at all frequencies, as the Nyquist stability condition requires. A

Fig. 9 Photo of the experimental platform

Fig. 10 Experimental results obtained using controller Re(s) for m 3.2 kg a Experimental tip angle ut and motor angle um obtained
b Comparison between the experimental and simulated motor voltage V
1458
IET Control Theory Appl., Vol. 1, No. 5, September 2007

Fig. 11

Experimental results obtained using controller Re(s) for m 1.9 kg

a Experimental tip angle ut and motor angle um obtained b Comparison between the experimental and simulated motor voltage V

Fig. 12

Experimental results obtained using controller Re(s) for m 0.6 kg

a Experimental tip angle ut and motor angle um obtained b Comparison between the experimental and simulated motor voltage V

Remark: Conditions (32) (34) make the closed-loop system stable for any single-link exible arm that fulls the interlacing property (31), independent of the number of high-frequency modes considered. 6 Experimental results

The control strategy proposed here, with the use of the outer loop controller in (30), has been tested in the experimental platform of the picture in Fig. 9, whose dynamics corresponds to the one described previously for the single-link exible manipulator. In this section, the experimental results obtained are presented. The robustness of the system has been tested by changing the payload at the tip. Motor and tip position records are shown. Simulated control signals are plotted together with the experimental motor control signals for comparison purposes. Note that the simulations neglect Coulomb friction, whereas in the experimental platform a compensation term, 0.3 V for positive motor velocities and 2 0.25 V for negative ones, has been added to the control signal. These compensation values have been found by a trial-and-error process. Fig. 10a shows the measurements of the tip angle ut and the motor angle um , whereas Fig. 10b shows the motor voltage V, both gures corresponding to a mass m 3.2 kg. A relay type control appears in the transient and steady states due to Coulomb friction compensation.
IET Control Theory Appl., Vol. 1, No. 5, September 2007

For this reason, the experimental voltage signal obtained presents quick oscillations and is not zero in the steady state. Fig. 11 shows the measurements of the tip and motor angles and motor voltage obtained for the nominal mass m 1.9 kg. And nally, Fig. 12 shows the results when m 0.6 kg. Through Figs. 10b, 11b and 12b, it can be observed that the peak of the control signal keeps lower than the saturation limit and remains almost constant in the presence of payload changes, with a value around 0.65 V, making this control strategy very suitable for motor saturation problems. Another important aspect to remark is that the fractional derivator part of the controller, s 0.85, is implemented by s 0.85 s(1/s 0.15). That is, the fractional integrator part acts like a low-pass lter of the signal that enters the derivative operator and reduces the noise introduced through the control loop. Therefore with the fractional controller, the system is not only more robust to payload changes, but also to noise presence. 7 Conclusions

A new method to control single-link lightweight exible arms in the presence of payload changes has been presented in this work. The overall controller consists of three nested control loops. Once the Coulomb friction and the motorbeam coupling torque have been compensated, the inner loop is designed to give a fast motor response. The
1459

simplifying loop reduces the system transfer function to a double integrator. The fractional order derivative controller is used to shape the outer loop into the form of a fractional order integrator. The result is an open-loop constant-phase system whose closed loop responses to a step command exhibit constant overshoot, independent of variations in the load. A study of the effect of spillover has been carried out, where system stability to any non-modelled higher-order dynamics has been proven. The fractional order controller has been tested in an experimental platform by using discrete implementations. From the results obtained, it can be concluded that an interesting feature of the fractional control scheme is that the overshoot is independent of the tip mass. This allows a constant safety zone to be delimited for any given placement task of the arm, independent of the load being carried, thereby making it easier to plan collision avoidance. It must be remarked that with the fractional order controller, the control signal is less noisy than with a standard PD controller, as the fractional integrator acts like a low-pass lter and reduces the effects of the noise introduced in the control loop. Besides, with this control strategy, changes in the payload imply only slight variations in the maximum value of the control signal, avoiding possible saturation issues. 8 Acknowledgments

This work was nancially supported by the Spanish Government Research Program via Project DPI-2003-03326 (M.E.C.) and by the Spanish Research Grant 2PR02A024 of the Junta de Extremadura. 9 References

1 Zinn, M., Khatib, O., Roth, B., and Salisbury, J.K.: Playing it safe [human-friendly robots], IEEE Robot. Autom. Mag., 2004, 11, (2), pp. 1221 2 Cannon, R.H., and Schmitz, E.: Initial Experiments on the endpoint control of a exible one-link robot, Int. J. Robot. Res., 1984, 3, (3), pp. 6275 3 Chen, Y.P.: Regulation and vibration control of a FEM-based single-link exible arm using sliding-mode theory, J. Vib. Control, 2001, 7, (5), pp. 741 752 4 Su, Z., and Khorasani, K.: A neural-network-based controller for a single-link exible manipulator using the inverse dynamics approach, IEEE Trans. Ind. Electron., 2001, 48, (6), pp. 10741086 5 Benosman, A., and Le Vey, G.: Control of exible manipulators: a survey, Robotica, 2004, 22, (5), pp. 533 545

6 Feliu, V., and Ramos, F.: Strain gauge based control of single-link very lightweight exible robots to payload changes, Mechatronics, 2005, 15, (5), pp. 547 571 7 Feliu, J.J., Feliu, V., and Cerrade, C.: Load adaptive control of single-link exible arms based on a new modeling technique, IEEE Trans. Robot. Autom., 1999, 15, (5), pp. 793804 8 Podlubany, I.: Fractional differential equations (Academic Press, San Diego (1999)) 9 Feliu, V., Vinagre, B.M., and Monje, C.A.: Fractional control of a single-link felexible manipulator. Proc. ASME IDETCT/CIE, , Long Beach, CA, 2005 10 Oustaloup, A.: The CRONE approach: theoretical developments and major applications. Proc. Second IFAC Workshop on Fractional Differentiation and its Applications, , Porto, Portugal, 2006, pp. 3969 11 Chen, Y.Q.: Ubiquitous fractional order control?. Proc. Second IFAC Workshop on Fractional Derivatives and its Applications (FDA06), 2006, Porto, Portugal 12 Le Mehaute , A., Tenreiro, J.A., Trigeasson, J.C., and Sabatier, J.: Fractional differentiation and its applications (Ubooks, 2005) 13 Feliu, V., Rattan, K.S., and Brown, H.B.: Control of exible arms with friction in the joints, IEEE Trans. Robot. Autom., 1993, 9, (4), pp. 467475 14 Kokotovic, P.V.: Applications of Asingular perturbation techniques to control-problems, SIAM Rev., 1984, 26, (4), pp. 501 550 15 Siciliano, B., and Book, W.: A singular perturation approach to control of lightwieght exible manipulators., Int . J. Robot. Res., 1988, 7, (4), pp. 7990 16 Karimi, H.R., Yazdanpanah, M.K., Patel, R.V., and Khorasani, K.: Modelling and control of linear two-time scale systems: applied to single-link exible manipulator, J. Int. Robot. Syst., 2006, 45, pp. 235265 17 Monje, C.A., Caldero n, A.J., Vinagre, B.M., Chen, Y.Q., and Feliu, V.: On fractional PIl controllers: some tuning rules for robustness to plant uncertainties, Nonlinear Dyn., 2004, 38, (1 4), pp. 369381 18 Vale rio, D., and Sa Da Costa, J.: ZieglerNichols type tuning rules for fractional PID controllers. ASME Int. Design Engineering Technical Conf. and Computer and Information in Engineering Conf, , Long Beach, CA, USA, 2005 19 Bode, H.W.: Relations between attenuation and phase in feedback amplier design, Bell Syst. Tech. J., 1940, 19, pp. 421 454 20 Oustaloup, A.: La Commade CRONE: Commande Robuste dOrdre Non Entier (Hermes, Paris, 1991) 21 Oustaloup, A., Levron, F., Nanot, F., and Mathieu, B.: Frequency-band complex noninteger differentiator: characterization and synthesis, IEEE Trans. Circuits Syst. I Fundam. Theroy Appl., 2000, 47, (1), pp. 25 40 22 Vale rio, D.: Fractional robust system control. Instituto Superior Te cnico, Universidade Te cnica de Lisboa, 2005 23 Podlubny, I., Petra s , I., Vinagre, B.M., OLeary, P., and Dorc a k, L.: Analogue realizations of fractional-order controllers, Nonlinear Dyn., 2002, 29, (1 4), pp. 281296 24 Chen, Y.Q., and Moore, K.L.: Discretization schemes for fractional-order differentiators and integrators, IEEE Trans. Circuits Syst. I Fundam. Theory Appl., 2002, 49, (3), pp. 363367

1460

IET Control Theory Appl., Vol. 1, No. 5, September 2007

You might also like