You are on page 1of 136

INFORMATION TO USERS

This manuscript has been reproduced from the microfilm master. UMI films the
text directly from the original or copy submitted. Thus, some thesis and
dissertation copies are in typewriter face, while others may be from any type of
computer printer.

The quality o f this reproduction is dependent upon the quality o f the copy
submitted. Broken or indistinct print, colored or poor quality illustrations and
photographs, print bleedthrough, substandard margins, and improper alignment
can adversely affect reproduction.

In the unlikely event that th e author did not send UMI a complete manuscript and
there are missing pages, these will be noted. Also, if unauthorized copyright
material had to be removed, a note will indicate the deletion.

Oversize materials (e.g., maps, drawings, charts) are reproduced by sectioning


the original, beginning at the upper left-hand com er and continuing from left to
right in equal sections with small overlaps. Each original is also photographed in
one exposure and is included in reduced form at the back of the book.

Photographs included in the original manuscript have been reproduced


xerographically in this copy. Higher quality 6” x 9” black and white photographic
prints are available for any photographs or illustrations appearing in this copy for
an additional charge. Contact UMI directly to order.

Bell & Howell Information and Learning


300 North Zeeb Road, Ann Arbor, Ml 48106-1346 USA
800-521-0600

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
The Waterbed Effect and Stability in Learning/Repetitive Control

Thuanthong Songchon

Submitted in partial fulfillment o f the


Requirements for the degree o f
Doctor of Philosophy
in the Graduate School o f Arts and Sciences

COLUMBIA UNIVERSITY

2001

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
UMI Number: 3005799

__ ___ __®

UMI
UMI Microform 3005799
Copyright 2001 by Bell & Howell Information and Learning Company.
All rights reserved. This microform edition is protected against
unauthorized copying under Title 17, United States Code.

Bell & Howell Information and Learning Company


300 North Zeeb Road
P.O. Box 1346
Ann Arbor, Ml 48106-1346

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
©2001

Thuanthong Songchoii
All Right Reserved

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
A bstract

T he W aterbed E ffect and Stability in L earning/R epetitive C ontrol

Thuanthong Songchon

The waterbed effect is a fundamental limitation to the performance o f

feedback control systems. It says that if the frequency response o f a feedback control

system has the property that it can significantly attenuate the effects o f disturbances

in one frequency range, then it must amplify disturbances in some other frequency

range. Learning and repetitive control often aim to eliminate all periodic disturbances

o f a fixed period, i.e. for a fundamental frequency and all harmonics. This suggests

that one must pay for this elimination by amplifying errors that occur between these

evenly spaced frequencies. In this thesis, learning and repetitive control are studied

using linear phase lead as well as low pass filtering for a frequency cutoff o f the

learning for purposes o f stabilization. It is shown that learning control has the ability

to bypass the waterbed effect, and the range o f mechanisms for doing so are

described. Technically, in finite time problems such as learning control one is never

in steady state response. Hence, the waterbed conclusions apply to that part o f the

trajectory beyond a settling time o f the system. In the repetitive control problem, time

progresses indefinitely, allowing the system to converge to steady state response. It is

shown that for real time repetitive control, one is subject to the waterbed effect. On

the other hand, it is possible to bypass this effect by making batch updates of the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
repetitive control signal, at least for periodic inputs and disturbances o f the period

being addressed. Sufficiently slow updates will result in the situation that applies to

learning control. Methods are developed to predict the disturbance to error steady

state frequency response characteristics for batch update repetitive control.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
T ab le o f C o n ten ts

Chapter 1 1
Introduction
1.1 Background 1
1.2 Dissertation Outline 2

Chapter 2 5
Stability Conditions of Learning and Repetitive Control
2.1 Introduction 5
2.2 A Frequency Response Based Stability Condition 6
2.3 The True Stability Boundary for ILC and RC 8
2.4 Stability o f First Order Systems 10
2.5 Stability o f Second Order Systems 13
2.6 Stability o f Third Order Systems 15
2.7 Conclusions 16

Chapter 3 23
Design Trade-offs in Learning/Repetitive Controllers Using Zero-
Phase Filtering

3.1 Introduction 23
3.2 Sources o f Tracking Error in Feedback Control Systems 23
3.3 Requiring Good Learning Transients in Learning and 25
Repetitive Control
3.3.1 Learning Control 25
3.3.2 The Good Transient Condition in Learning Control 28
3.3.3 Repetitive Control 30
3.4 An Example System for Numerical Investigations 32
3.5 What Do We assume We Know About The System When 33
We Design The Learning or Repetitive Controller
3.6 The Design Approach 34
3.6.1 Compensator Design 35
3.6.2 Zero-Phase Low Pass Filtering 36
3.6.3 Purpose o f This Chapter 39
3.7 The Steady State Error Level When Filtering The Total Command 39
Signal Going Into The Feedback Controller
3.7.1 Iterative Learning Control 39
3.7.1.1 A Simple Special Case 41
3.7.1.2 Numerical Example Using a Zero-Phase 42
Butterworth Filter

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
3.7.2 Repetitive Control 43
3.7.3 Periodic filtering in Learning Control 45
3.8 The Steady state Error level W hen The Desired Trajectory Part 45
o f The Command Is N ot Filtered
3.8.1 Iterative Learning Control 45
3.8.2 Comparison W ith the Previous Case 47
3.8.3 Repetitive Control 48
3.9 Conclusions 49

Chapter 4 57
The Waterbed Effect and Iterative Learning Control Using Zero-
Phase Filtering

4.1 Introduction 57
4.2 The Standard W aterbed 60
4.3 A Robot Example 62
4.4 Generalized Bode Integral Theorem 62
4.5 Finite Time Trajectories and The Frequencies in GBI Theorem 66
4.6 Implications for Causal ILC Laws 66
4.7 GBI Theorem and N oncausal ILC 68
4.8 Low-Pass Filter Learning C utoff 70
4.9 GBI Theorem and Zero-Phase Filtering 72
4.10 Conclusions 74

Chapter 5 11
The Waterbed Effect and Repetitive Control with Filtering
5.1 Introduction 77
5.2 Iterative Learning Control and Repetitive Control Background 78
5.3 Batch Zero-Phase Low-Pass Filtering 81
5.4 Real Time Low Pass Filtering 85
5.5 Some Waterbed Conclusions 86
5.6 The Need for Zero-Phase Filtering in RC 92
5.7 Understanding The M echanism for Amplification 93
5.8 Comments on Stability Conclusions and Stability Boundary 95
5.9 Frequency Response o f Periodic Update Repetitive Control 99
5.10 Conclusions 110

Chapter 6 121
Conclusions

Refferences 123

ii

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
A c k n o w le d g m e n ts

I would like to express my deep appreciation to Professor Richard W.

Longman for his support, encouragement, and guidance w hile working on this

dissertation, and for his kindness throughout the doctoral program.

I w ould like to acknowledge my parents, Mr. Prachoke Songchon and Mrs.

Pleum jitr Songchon, for their constant love, support, and education. Also, I would

like to thank my sisters, M s.Krisana Songchon and M s.Bavom nan Songchon for their

love and support.

My warm est thanks m ust go to my family, M rs.M onrawee Songchon and my

son. Titisun Songchon. for their patience and love.

iii

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
1

C hapter 1

Introduction

1.1 Background

One major objective o f automatic control is tracking a desired trajectory. The

fields o f learning and repetitive control, apply to situations in which the same

command is given to a feedback control system many tim es, or the command is

periodic, and these controllers examine the tracking error o f the system from each

repetition or period, and adjust the command the next repetition or period in order to

try to make the error go to zero. Thus, by simply changing the com m and given to a

feedback controller, these methods can substantially increase the accuracy of an

existing feedback control system. An important special case o f repetitive control is a

control system that has a desired output that is constant, but it is subject to periodic

disturbances.

Iterative learning control (ILC) refers to methods o f iteratively adjusting the

command to a closed loop control system, to converge on that command which

produces zero tracking error following a desired trajectory. The system is restarted

from the same initial condition each time a command is given. Repetitive control

(RC) applies to systems with a periodic desired output, or with a constant desired

output and there are periodic disturbances, and the command to the closed loop

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
system is adjusted from one period to the next in order to converge to zero tracking

error.

It is interesting to note that the concept o f iterative learning control (ILC)

appeared simultaneously in the early 1980s. Arimoto et al. [1]. Casalino and Bartolini

[4] and Craig [5] are independent developments o f similar ideas that year, with

Uchiyama [35] being one o f the precursors. Middleton et al. [23], subm itted in 1984.

was another independent developm ent motivated by robotics but using repetitive

control. The origin o f repetitive control (RC) had different motivation, and early

works include Inoue et al. [16], O m arta et al. [26], Hara et al. [10,11], Nakano and

Hara [24], and Tomizuka et al. [34], These two fields appear to be very different in

the literature, but we show that in practical use they are not so different.

This dissertation investigates the stability o f learning and repetitive control,

develops methods to stabilize learning and repetitive control by zero-phase filtering,

and develops an understanding o f the relationships between the waterbed effect and

learning and repetitive control.

1.2 Dissertation Outline

The outline o f the dissertation is as follows. Chapter 2 investigates conditions

for stability for both learning and repetitive control. In learning control the stability

boundary is presented as given in Phan and Longman [27], For repetitive control the

methods are different. One condition for defining the true stability boundary, makes

use o f the method o f Pierre [28], and the second condition is an approximate

frequency response based monotonic decay condition o f Elci et al. [7], In this chapter.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
these stability conditions are compared for first order, second order, and third order

systems for both learning and repetitive control, [31]. It is shown that for practical

applications there is little difference between these conditions in repetitive control.

Chapter 3 discusses a m ethod to stabilize learning and repetitive control by

using a zero-phase filter. In previous works o f Elci et al.[6,7] and H sin et al.[l3],

zero-phase infinite impulse response (HR) low-pass filtering is used. In this chapter

we develop m ethods o f predicting the steady state error levels that will be reached for

each choice m ade by the control system designer, [21].

Chapter 4 formulates methods to analyze the waterbed effect in learning

control. The original work o f Bode [2] used analytic function theory to examine the

properties o f a continuous feedback loop in the frequency domain. Later, many others

such as, in Seron et al. [30], studied the analogous results for the discrete time case.

The discrete time version is o f interest in here. All the methods are restricted to the

case o f all poles o f the sensitivity transfer function being inside the unit circle, and the

sensitivity transfer function being a proper rational function. In learning and repetitive

control, improper sensitivity functions often occur, especially when one uses zero-

phase filtering. In this chapter we extend the Bode integral theorem to the more

general cases needed for learning and repetitive control systems [32].

Chapter 5 studies the waterbed effect and repetitive control using filtering for

a frequency cut o ff o f the learning for purposes o f stabilization. It is show n that for

real-time repetitive control, one is subject to the waterbed effect. On the other hand, it

is possible to bypass this effect for the frequencies being learned by m aking batch

updates o f the repetitive control signal. Sufficiently slow batch updates in repetitive

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
4

control results in the same situation that applies to iterative learning control for the

frequencies being learned, where it is possible for certain classes o f learning control

laws to bypass the w aterbed effect. Methods are developed to predict the disturbance

to error frequency response characteristics for the range o f possible batch update rates

[33].

Chapter 6 gives conclusions.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
5

Chapter 2

Stability C onditions o f Learning and R epetitive Control

2.1 Introduction

Stability is one o f the major concepts in control systems. Without stability, the

goal o f the control cannot be achieved; with stability the system can at least be used.

In both iterative learning control (ILC)-and repetitive control (RC) cases it is common

to aim to satisfy a frequency response based stability condition, normally heuristically

derived to suggest convergence based on decay o f the steady state frequency response

components o f error. Elci et al. [7] and Longman [19] show that this condition is a

sufficient condition for asymptotic stability o f the learning process, but suggests that

its real use is as a technique for ensuring good transients during the learning process.

Huang and Longman [15] suggest that the discrepancy between this frequency

response based condition and the true stability boundary is very large for ILC, but for

RC the condition will normally be very close to the true stability boundary. It is the

purpose of this chapter to investigate the distinction between the frequency response

based condition and the true stability boundary, for the set o f all first order systems,

all second order systems, and all third order systems with no zero. This is done for the

most basic form o f ILC and RC, integral control based learning. In the case o f ILC

this adjusts the command at time step k o f repetition j according to Uj-i(kT) = Uj(kT)

+<fiej((k+l)T), where <p is the learning gain, T is the time step interval, and e, is the

tracking error (desired output minus actual). The one step time shift in the error

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
6

accounts for the usual one step delay between a change in the input to a change in the

output in a digital system . In repetitive control the change in repetition num ber j is

replaced by a tim e shift o f p time steps corresponding to one period. W ritten in terms

o f z-transform s this becomes 2?U(z) = U{z) + (jrzE(z). The associated repetitive control

system block diagram has the desired output as the command, uses unity feedback,

has a repetitive controller box with transfer function <f/zl(zp - I ) , and the plant is the

transfer function o f the closed loop feedback control systems w hose input is being

adjusted, and whose output is to converge to the desired output.

2.2 A Frequency Response Based Stability Condition —A n Approxim ate

M onotonic Decay Condition

A com m on stability condition for both iterative learning control and repetitive

control asks that

( 2 . 1)

for all frequencies co up to Nyquist. where T is the sample tim e interval, and Gc(z) is

the z-transfer function o f the associated closed loop control system . It can be shown

that this condition is a sufficient condition for convergence to zero tracking error for

both ILC and RC (Longm an [19], Elci et al. [7], Huang and Longm an [15]). These

references suggest that its real importance is as an approximate condition for assuring

monotonic decay o f the tracking error with repetitions or periods.

To see this for ILC, suppose that the output o f the closed loop control system

is Y{z) = Gc{z) + W(z) where W represents any disturbance that appears every time the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
7

command is given. Write this for repetitions /+ 1 and j, and take the difference.

Express the result in terms o f a difference o f errors, and use the learning control law-

described in the previous section to produce the input in terms o f the error. The result

can be w ritten as Ej+\(z) = [1 -<fizGc(z)]Ej(z) (note that the initial condition on E is zero

for ILC). By satisfying (2.1). it is guaranteed that the steady state frequency response

components o f the error will decay monotonically with repetitions. However, since

the trajectory is a finite time trajectory, technically it is never in steady state.

Nevertheless, w hen the trajectory is significantly longer than a few time constants o f

the system, this condition makes a good condition to satisfy in order to ensure good

learning transients, by creating monotonic decay o f the error.

For RC, find the transfer function from the periodic desired trajectory' YJ^z) to

the associated error, for the block diagram described in the previous section. This

produces [ z p - 1 + <pzGc(z)\E(z) = ( z p - 1 )[Yl{( z ) - I V ( z ) ] . The right hand side is zero

due to the periodicity with period p o f the desired trajectory and the disturbance. This

makes a homogeneous difference equation whose transients determine the

convergence o f the error. Rewrite this equation as z pE(z) = [1 -tpzGc(z)]E(z) and

note that the multiplication by z p is a shift one period forward in time. This suggests

that if (2.1) is satisfied, there will again be monotonic decay o f each frequency

component o f the error. This time there is a quasi-static assumption made, in order to

have steady state frequency response thinking apply.

If one chooses to satisfy (2.1) in order to obtain good transient o f the learning

process, it is o f interest to know how much more restrictive satisfying (2.1) is than

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
8

simply satisfying the if and only if condition for stability. In this paper we show how

these differ for ILC and RC for first, second, and third order systems.

2.3 The True Stability Boundary for ILC and RC

In ILC the true stability boundary when using integral control based learning

with learning gain ^ , is given by

0 < {CB)<f> < 2 (2.2)

where B and C are from the discrete-time state-space model x((/t+l)7~) = Ax(kT) +

Bu(kT). y{kT) = Cx{kT) o f the closed-loop single-input, single-output (SISO) system

(see e.g.. Phan and Longman [28]).

In repetitive control, the most natural way to express the true stability

boundary is in terms o f the Nyquist stability criterion. The repetitive control loop

contains the closed-loop feedback control system Gc( z ) and the repetitive control

law <pzl(zp - \ ) together with unity feedback, so that the characteristic polynomial

can be written in the form l-i-<f>Gr(z) = 0 with Gr(z) = z G c( z ) / ( z p - 1 ) . It will be

convenient for us to separate the DC g a i n o f the feedback control system from the

rest o f its transfer function according to Gc (z) = KcG ( z ) , and then define the product

o f the DC gain with the learning gain as K = Kc<f>, and then this serves as our gain

parameter. The characteristic equation becomes 1 + KG(z) = 0. Direct application o f

the discrete time Nyquist criterion is inconvenient because o f the p roots on the unit

circle. Following Huang and Longman [15], we apply the method o f Pierre [28] to

handle this difficulty. Rewrite the characteristic polynomial in the form

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
9

-K + 0 (z) = 0, Q(z) = -1 !G(z) (2.3)

The troublesome poles on the unit circle become zeros o f O, simplifying the plotting.

Plot Q{e ' ° ) for 0 going from 0 to 180 deg, deleting any points for which O is singular

(one does not have to go around these points). For any gain K as a point on the real

axis o f the O plane, where lm 0 ( e '° ) ^ 0 , then

Z = ( - W /180°) + Pq 4 - ( « ’/ 2) (2.4)

where

W is the angle swept by the vector pointing from point (AT, 0) to the moving

point O(e ‘0) for 6 going from 0 to 180 deg with singularities deleted. Clockwise is

counted as positive.

Pq is the total number o f poles o f O outside the unit circle, finite poles plus

poles at infinity.

Z is the number o f zeros o f -K + Q(z) = 0 that are outside the unit circle.

ri is the number o f poles o f O on the unit circle.

In using (2.4), one normally knows the values for Pq and r i , W is determined from

the plot, and hence Z is known. The system is stable for all K that produce Z equal

zero. Note that this stability condition which represents the true stability boundary,

depends on the number o f time steps p in a period, whereas the previous approximate

monotonic decay condition (2.1) does not.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
10

2.4 Stability o f First O rder Systems

The Stability Conditions: N ow let us examine the distinction between stability

condition (2.1) and the true stability boundary (equation (2.2) for ILC and (2.4) for

RC) for all possible first order systems. Start with a continuous tim e transfer function

G v(s) = K ca /(s + a) where K c is the DC gain. When fed by a zero order hold, this

converts according to the rule Gc(z) = (\ —z~x)Z\Gs( s ) l s \ where the Z indicates to

take the z-transform o f the function represented in the square bracket. Then the Gc(z)

for equation (2.1) is

Gc(z) = K c(l —e ~aT)/( z —e ~aT) (2.5)

The G(z) for equations (2.3) and (2.4) is given by

G (z) = (l —e~aT)z/[(z —e ' aT) ( z p —1)] (2.6)

And condition (2.2) becomes

0 < <pKc(1 —e ' a‘ ) < 2 (2.7)

There are three parameters whose values may affect these stability conditions: the

gain K = <pKc which is the product o f the learning gain with the DC gain o f the

system, the value o f aT related to the time constant o f the system and the sample time,

and the num ber o f time steps p in the desired trajectory or period.

Concerning Limits on the Parameters: In order for a discrete tim e control system to

function well,one should have the sample rate be such that there are several time

steps in a tim e constant o f the system. In this case the time constant is1/a. so a

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
11

generous upper limit on the value o f a T is unity. W hen we get to second order

system s s 2 + 2^a>0s + col - we need at least one sample per time constant when the

roots are real, and when the roots are complex we again ask for one sam ple per time

constant for the real part o f the root, and at least two sample times per period for the

oscillatory part o f the root (this lim its values o f co0T to a maximum o f n r / -Jl - C 2 ).

The Approximate Monotonic Decay Condition: The range o f gain K satisfying

condition (2.1) can be found by plotting <pe'°Gc(e‘e ) for £?going from 0 to 180 deg,

and seeing how large K can be before the curve goes outside the unit circle centered

at +1. This happens first w hen 9 = 0 . and produces the inequality 0 < K < 2.

Condition (2.1) is always independent o f p , but in this case it is also independent of

param eter a T as well. As stated above, this is a sufficient condition for stability for

both ILC and RC.

The True Stability Boundary f o r Learning Control: The discrete tim e state variable

representation o f (2.5) has matrices A, 5, C given by e~“r . K c( \ - e ~ al )A

respectively. Then the stability boundary is given by 0 < K < 2/(1 —e~a l). This is

independent o f p, and the stable range on K tends to infinity as the sam ple time tends

to zero. The boundary is show n in Fig. 2.1. It is always larger than the sufficient

condition (2.1) as it must be, and it becomes arbitrarily larger as the sample time

approaches zero.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
12

The True Stability Boundary f o r Repetitive Control: For conciseness, denote sine by

s and cosine by c. Then the 0 ( e 'e ) from (2.6) can be written as

0 ( e '° ) = (1 - e-"r ) '[ 1 - c ( p 0) + e-aT(c((p - 1 )0 ) - c 0 )


+ i(e~aT(s((p - 1 ) 0 ) + s0) - s { p 0 )]

A typical plot o f Q(e'°) is given in Fig. 2.2, p — 10 and a T = 0.88. Applying the

modified Nyquist plot rules above to any point (K.0) between 0 and the first time the

plot crosses the positive real axis produces, W = 5 • 180°, Pq — 10, n' = 0 with Z = 0.

Hence, all gains K between zero and this first crossing o f the positive real axis

correspond to stability. For this first order system, it happens that the first loop is

always the one determining stability, but for the second and third order systems

discussed below, this is not necessarily the case. Then, the procedure for determining

the maximum stable gain Kmax as the parameters o f the system are run through their

range o f values, is:

(i) Set the imaginary part o f 0 ( e ' ° ) equal to zero. For the case o f equation (2.8)

this can be rewritten as

e~aTc{{p - 2 ) 9 / 2 ) - c ( p 9 / 2 ) = 0 (2.9)

Then solve this numerically to get the p solutions for 9.

(ii) Substitute each solution for 9 into O (ei0) to find the associated value o f K

according to (2.3). The minimum o f these values is the gain for the stability

boundary, K max.

Figure 2.3 shows the results o f this procedure for the first order system, giving

Kmax for a T in the range o f reasonable values from 0 to I, and for various values o f

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
13

the number o f tim e steps in a period, p. As m ust be the case, the stability boundary is

always above the value K = 2 given by the monotonic decay condition (2.1).

However, this difference is only substantial w hen the num ber o f time steps in a period

is quite sm all, e.g. for a p o f 10. In typical digital control systems with sam ple rates

like 100 or 1000 Hz, any reasonable length period for the desired m otion will have a

p sufficiently large that the distinction between the true stability boundary in

repetitive control, equation (2.4), and the approximate monotonic decay condition

(2.1), becom es insignificant. Hence, in most practical situations, satisfying condition

(2.1) is close to a requirement, even though it does not correspond to the true stability

boundary. T he difference between (2.1) and (2.4) is insignificant for typical length

trajectories. A nd use o f (2.1) in designing repetitive controllers is much easier than

using (2.4).

2.5 Stability of Second Order Systems

N ow consider the set o f all stable strictly proper second order system s. The

transfer function in continuous time is G f(s) = K cco] (ds + l ) / 0 r + 2Ccoos -hco2) . and

we consider th at it is fed by a zero order hold, and then convert to the associated z-

transfer function. This time the parameters that can influence stability are K , coaT ,C ,

d/T. and p .

For underdam ped systems ( 0 < C < 1)

Gc(z) = K c(A,z + B , ) / ( z 2 - 2 e- a' c f r z + e~2a' )

A, =1 - e ~ a' c f t - { c c x/ p x)e-a's(3x + ( r 2S / / 3 x)s/3x

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
14

A = <T2"' + ( a , / f r ) e - a'sfr —e~a'c/3l - ( r 2^ / y f f ,M

a, = ^ , T - A = c o j 4 \ - C 1 , r = cotJ , S = d !T (2.10)

For overdamped systems ( t ' > 1)

G c (z) = K c (A2z + 5 ,) /[(a , - A ) ( ^ A r 2 - ( e ^ + )- + D]

.4 , = ( A ^ - l ) a 2e “: + (1- » , £ ) & « * + ( a , - P 2 ) e a '-+ p '-

B2 = ( f 2S - l ) a 2e fi- + ( \ - a 25)jB2e a'- + a 2 - / 3 2

a 2 =( o 0t {c + 4 ^ - - \ \ J32 = c o j { c - 4 C - 1) (2-11)

For critically damped systems

Gc(z) = £ c( ^ z + fl3) /(e 2rz 2 - 2 e rz +1)

A3 = ( r 2J - r - l ) e r + e 2r , A = 1+ ( r - r 2^ - l)er (2.12)

These Gc(z) are used in (2.1) for the approximate monotonic decay condition.

The conversion o f Gc(z) to G(z) for use in (2.4) is analogous to equations (2.5) and

( 2 .6 ).

The True Stability Boundary f o r Learning Control: For the Gc(z) above for the

underdamped, overdam ped and critically damped cases, we can convert the

associated second order scalar difference equation into state variable form, and

substitute into the ILC stability condition (2.2) to obtain, respectively,

0<K<2/Al

0 < K < 2 ( a 2 -J32)ea'-+p'-1 A, (2.13)

0 < K < 2l(A5e'laiJ)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
15

The results are shown in Fig. 2.4 for the case d = 0, where again the range on K goes

to infinity as the sample tim e T goes to zero.

The Approximate Monotonic Decay Condition: It is not possible for second order

systems to obtain a simple analytical expression for the range o f K as was done for

the first order case. Figures 2.5 and 2.6 give the results for the cases o f d!T = 0 and

d/T = 1 respectively. Again K is lim ited by 2, the diameter o f the unit circle the plot

o f (pzGfz) should stay within. But this time it is not only along the real axis that one

might start leaving this circle when the gain is too large, but when there is a resonant

peak it can easily leave earlier at some frequency other than zero. Hence, as C

decreases, the range o f K is made smaller and smaller. All results are, o f course,

independent o f p , and the monotonic decay condition is vastly different than the true

stability boundary in the case o f ILC.

The True Stability Boundary f o r Repetitive Control: Figures 2.7 and 2.8 give the

true stability boundary using equation (2.4). It is seen that for overdamped systems

with a very small number o f time steps p per period, it is possible to go slightly above

the K equal 2 value. Under other circumstances, the distinction between the

monotonic decay condition (2.1) and the true stability boundary is very slight.

2.6 Stability o f Third Order Systems

The same procedure is applied to third order systems o f the form

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
16

f 1 \
Gx{s) = K cf d 1 (2.14)
+ d y U 2 +2£a>0s + a ) ; )

After discretizing, the parameters are K, conT , dT, C . and p. The equations involved

are much more complicated than those o f the second order system and are not shown.

Figures 2.9 and 2.10 give the monotonic decay condition (2.1) results, and Figures

2.11 and 2.12 give the true stability boundary. The implications o f these plots are

similar to the second order case.

2.7 Conclusions

This chapter has shown that the approximate monotonic decay condition (2.1).

a sufficient condition for stability, is sufficiently close to the stability boundary for

repetitive control that in practical applications one should aim to satisfy it. Condition

(2.1) differs from the true stability boundary substantially for very small p . and for

first order systems, and otherwise the difference becomes negligible. As discussed in

Longman [19] it is important to satisfy condition (2.1) for ILC in order to obtain

reasonable learning transients, in spite o f the very big difference between (2.1) and

the true stability boundary in the ILC case.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
17

20

0.2 0.4 0.8


aT

Fig. 2.1 The stability boundary o f learning control applied to the first order system.

a.

real part

Fig. 2.2 Plot o f 0(e 0), with # fro m 0-180°, o f the first order system w hen p =10 and

aT = 0.88.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Fig. 2.3 True stability boundary o f repetitive control applied to the first order system.

450

400

350

300

250

^ 200

150 = 0.2

100

= 0.2

Fig. 2.4 The stability boundary o f learning control applied to the second order system

when d /T = 0 and <£"= 0.2, 0.5, 0.7, 1. and 2 from left to right.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
19

0.7

= 0.5

0.5 = 0.2

Fig. 2.5 Monotonic decay boundary o f repetitive control applied to the second order

system when d/T = 0.

= 0.7

= o.

0.5
= 0.2

Fig. 2.6 Monotonic decay boundary o f repetitive control applied to the second order

system when d/T = 1.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
20

2.5

1.5
r = 0.5, p= 50
- ->m

0.5

Fig. 2.7 True stability boundary o f repetitive control applied to the second order

system w hen d/T = 0.

r = i,p = 5 0
s 200
A 10.
r= 0.7, p=50
* onn

0.5

Fig. 2.8 True stability boundary o f repetitive control applied to the second order

system w hen d/T = 1.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
21

£ = 0.2

0.5

Fig. 2.9 M onotonic decay boundary o f repetitive control applied to the third order

system when clT = 0.5.

= 0.5, /-= 0.7


O0 o
= 0.2

0.5

3.5

Fig. 2.10 Monotonic decay boundary o f repetitive control applied to the third order

system when dT = 1.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
22

2.5

/- = 0.2, p=50 r = 0.5, p= 50 r= 0.7, p=50


- 200 ’ ->nt *= onr
- o —V a

0.5 C= 2, p= 50
- 200

4.5

Fig. 2.11 True stability boundary o f repetitive control applied to the third order

system when dT = 0.5.

2.5

1.5

0.5

2.5 3.5 4.5

Fig. 2.12 True stability boundary o f repetitive control applied to the third order

system when dT = 1.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
23

C hapter 3

D esign Trade-offs in L earning/R epetitive C ontrollers Using Z ero-

Phase Filtering

3.1 Introduction

In learning and repetitive control, guaranteed mathem atical convergence to

zero tracking error is not enough. One needs to impose conditions to ensure good

transients during the learning process. A natural condition to impose is monotonic

decay o f the errors with repetitions o f the task, or with periods o f the periodic

command, and this can be done in the frequency domain. However, it is very difficult

to satisfy this monotonic decay condition for all frequencies. In previous work

summarized in [18], some simple learning and repetitive control design methods were

developed. They are similar to designing a PID controller in classical control,

requiring the adjustment o f a few parameters. They achieve the desired mono tonic

decay o f the errors, but do so at the expense o f not fixing the errors at all frequencies.

In this paper we develop m ethods for the learning or repetitive control system

designer to use in picking his learning law or the parameters o f his learning law, in

order to achieve both good learning transients, and good final values o f the error.

3.2 Sources of Tracking Error in Feedback Control Systems

Consider a typical digital feedback control system as in Fig. 3.1. This

particular form uses a digital input, and the feedback m easurem ent device gives a

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
24

digital output. The command input U(z) to a feedback control system is usually the

desired trajectory YJ^z). W e choose to look at the continuous time output only at the

sam ple tim es, so we include an A/D converter on the output. The K(s) is a

disturbance input appearing in the most com m on location, ju st before the plant

an d after the controller G[(z). It will be convenient to convert this disturbance to an

equivalent output disturbance. We will do it here for this typical digital control

system , but it can be done in an analogous w ay for m ost any more general system

w ith disturbances in any location. When a continuous tim e system Gjc(s) is fed by a

zero order hold with input M(z), the outputs at the sam ple times are given by using

the discrete time transfer function G 2(z) = (1 —z~x)Z(Glc( s ) / s ) , where the

Z ( G 2c(s) / s) represents the z-transform o f the time function corresponding to the

transform in parentheses. Superposition applies to include the influence o f the

disturbance on the output o f the plant block, but since the disturbance does not go

through a hold, we m ust solve the differential equation with the disturbance as its

input, and then add the result V(z) = Z(Gic{s) KG)) to the output, as in the top o f Fig.

3.2. From the block diagram, solving for Y(z) in terms o f the input U(z) and the

disturbance produces the equation

Y(z) = G(z)U(z) + W(z) (3.1)

V(z) (3.2)

N o w G(z) represents the closed loop discrete time system dynamics (closed loop

difference equation) and W(z) represents an output disturbance function.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
25

For normal feedback control, we make the com m and equal the desired output,

U{z) = YJjz), and then we see from equation (3.1) that the actual output Y(z) will

normally have error o f two kinds:

1. The inverse transform o f the first term in (3.1) is a convolution sum o f the inverse

transform s o f G(z) (the unit pulse response) and the desired output YJz). We can also

include initial conditions into this particular solution if w e wish. The convolution sum

o f a function with yd(k) will almost never equal yjjc), and hence there is a

determ inistic tracking error for almost all desired trajectories. We discuss the use o f

learning and repetitive control to try to eliminate this deterministic source o f error

2. The disturbance W{z) is a second source o f error. This disturbance can be from

random causes, and we will not concern ourselves with this type o f disturbance in this

paper. It can also be from deterministic sources. There are many situations in which

the same determ inistic disturbances occur every time you give a specific command.

An exam ple is the disturbance torque on a robot link as it follows a given path

through space. It is this class o f disturbance, written as an output disturbance, which

we wish to address by use o f learning and repetitive control.

3.3 R equiring Good Learning Transients in Learning and Repetitive Control

3.3.1 Learning Control

Learning control applies to problems in which a feedback controller is asked

to perform the same trajectory many times, always starting from the same initial

condition. Based upon the error observed in the previous repetition the command to

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
26

the feedback controller is adjusted in the next repetition to try to decrease the error.

To m ake a general formulation for linear learning control (as in [27]) we can convert

the discrete tim e closed loop system G(z) to m odem control form

xs(k+1) = AsXs(k) + Bsu(k), k = 0, 1, 2 ,..., p - 1

y s(k+\) = Csxs{k+\) +w {k) (3.3)

There is a p step desired trajectory, y j j c +1). The initial conditions each time are r(0).

and the same output disturbance w(k +1) appears every repetition. We assume that the

time delay o f the input-output map o f the digital feedback control system is one time

step. This is the expected delay for a digital control system, and when the delay is

different, one would usually know the value and adjust the mathematics accordingly.

The solution to (3.3) can be written as

y( k) = CsAskxsi 0) + j ^ C sAsk~'~lBsu(i) + w(k) ; k = 1,2,..., p (3.4)


/=0

Denote the variables for repetition j by using a subscript / , and write <5j in front of

any variable to form the difference: the value o f the variable at repetition j minus that

at repetition j - 1. Let y_. be a column vector o f the history o f the p outputs at

repetition j , starting with yj( 1) at the top. Let Uj be a column vector o f the

corresponding inputs starting with w/0) (note the one time step shift between these

two colum n vectors corresponding to the one step delay from discrete time input to

discrete tim e output). In terms o f these definitions, equation (3.4) can be written as

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
A general linear learning control law takes the form

8Jh. = LeJ-\'’ eJ = y J - y J (3-6)

In other words, the command to the feedback control system at repetition j is changed

from that at repetition j - 1 by adding a m atrix o f learning gains L tim es the error

observed in repetition j - I. We suppose that at repetition 0, before any learning has

taken place, we simply apply the desired output as the command to the feedback

controller. This means that the control history applied to repetition j + 1 is the sum

j
\LJ+x = l L j + L e J = y j + Y JLe l (3.7)
i=0

Reference [18] discusses the various form s o f the matrix L for different learning

laws. Noting that 8} y = - 8 j e . the error histories from one repetition to the next obey

e ^ \ = { I ~ P ,I ) e , (3.8)

and will converge to zero tracking error if and only if all eigenvalues o f U ~ are

less than unity in magnitude. It is relatively easy to satisfy this convergence

condition. For example, integral control based learning control uses a diagonal matrix

L. with the same number <p in each entry on the diagonal. The convergence condition

becomes |l - CsBs<j>\ < 1 which can be satisfied by a sufficiently small <f> o f the right

sign, and is independent o f the dynam ics o f the system in the matrix Asl Although

such a law has this very impressive stability robustness, being independent o f the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
28

system dynam ics, it can be impractical because o f poor transients [15]. Consider an

example. For the third order plant described in the exam ples section below, we use a

sample rate o f 100 Hz, and ask for a desired trajectory o f y d (0 = sin(2^r) on the time

interval from 0 to 1. A learning gain o f <f>= 1 is used. The result is shown in Fig. 3.3.

The root mean square o f the tracking error decreases from 0.4330 in repetition 0

w hen the desired trajectory is given as the command to the feedback controller, to a

m inim um at repetition 7 o f 0.1402. The mathematics above proves that the error will

converge to zero, but after repetition 7 the error starts to grow, reaching a maximum

RM S error level o f 1.1991 x 1051 at repetition 62,132. Then the error decreases and

ultimately reaches a numerical zero. For example, at repetition 3 x 103 the error has

reached 1.3145 x 10“*8. In order for the digital computer to be able to handle this

problem, we had to simplify it, by choosing a shorter trajectory, and using 100 Hz

instead o f 400 Hz as the sample rate. Otherwise, overflow occurs, and the iterations

stop before we can observe convergence.

3.3.2 The Good Transient Condition in Learning Control

Consider converting to the frequency domain, and examine the relationship

between the amplitudes o f the frequency components o f the error from one repetition

to the next. This means that we are characterizing the response in each repetition by

the steady state frequency response components. This assum ption is valid provided

that the trajectory is relatively long compared to a settling tim e o f the feedback

control system, so that the majority o f the trajectory is not influenced by transient

terms from the initial conditions.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
29

We will also restrict the choice o f the learning law L slightly, in order to make

it correspond to a time invariant law. This requires that all entries on the diagonal be

identical, all entries on the subdiagonal be identical, etc. And, as in the case o f the

initial condition transients, we assume that the matrix L is sufficiently large that edge

effects in the matrix can be ignored. Then, we convert the equations to the z-

transform domain, and to get the frequency response, we substitute z = e'mT where T

is the sample time, and co is the radian frequency under consideration. Then, the

equations analogous to (3.5), (3.6), and (3.8) become

SJY(euur) = - 8 JE{e,D,r) = G{e,a,r)SJU(e'a,T) (3.9)

S j U {e iaT) = L(e'ojr)Ej_i(e'QlT) (3.10)

E i (ei“T) = [ l - G ( e l“T)L(eiaT)]Ei .(e i“r )


(3 11^
= [1 - G(e'mT)L{e,air )]y E0(e,ojr)

Here the G(z) is the z-transfer function o f equation (3.1), and the L(z) is the

transform o f the learning law associated with each row o f L. For exam ple, L(z) = $rz

in the case o f the integral control based learning control. If the L were similar except

that the nonzero entries were on the super diagonal, shifted one entry to the right from

the true diagonal, then I ( z ) = ^k2, and if it is shifted by y —1 entries, then

L(z) = <pzr . When there are entries in more than one diagonal, the L(z) is formed

analogously, using superposition.

Equation (3.11) says that going from one repetition to the next, the amplitude

o f the component o f the error at frequency co is multiplied by \l —G ( e ,cur )L(e“u! ) |.

For any frequency for which this number is less than unity, the error will decay

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
30

monotonically, and any frequency for which the num ber is greater than unity, the

amplitude will grow m onotonically. Thus, if we require that

(3.12)

for all co up to the N yquist frequency, then all amplitudes o f the frequency

components o f the error decay monotonically. This statem ent applies to all portions

o f the trajectory for w hich steady state frequency response thinking can be applied.

This thinking suggests that (3.12) defines the stability boundary, but this is not the

case as explained in [15]. However, it is a sufficient condition for stability and

convergence to zero tracking error [6]. Nevertheless, when the desired trajectory is

sufficiently long com pared to the settling time o f the system that steady state

frequency response thinking can be applied, then satisfying (3.12) assures mono tonic

decay of all frequency components o f the error.

3.3.3 Repetitive Control

In repetitive control there is no restarting o f the system at repeating initial

conditions. Instead, the command is a periodic function o f time, and we wish to

eliminate tracking errors that appear each period. A special case o f particular

importance is when the command is a constant (periodic with any frequency) and one

wants to eliminate the effects o f a periodic disturbance w(k). We use the same form o f

the control law, except that repetitions become periods, where one period o f the

periodic command or disturbance has p time steps. The relationship of equation (3.10)

becomes

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
31

z pU{z) = U{z) + L{z)E{z) or U( z) = - ^ ^ E ( z ) (3.13)

The block diagram o f applying this repetitive control to the closed loop feedback

control system G{z) o f equation (3.1) is shown in Fig. 3.4.

From the block diagram we can solve for the error in terms o f the two inputs

to obtain

[ z p - 1 + L{z)G(z)]E(z) = ( z p - 1 )VYd { z ) - W { z ) ] (3.14)

A fter clearing fractions, this represents a difference equation. Note that the right hand

side represents the difference o f values shifted forward in time by p time steps minus

the current values. Since y j k ) and w(k) are both periodic with period p time steps,

this difference is zero and the right hand side o f (3.14) is zero. Therefore, the

difference equation is a homogeneous difference equation. Convergence to zero

tracking error occurs if all roots of the characteristic polynomial have magnitude less

than one.

As in the learning control case above, suppose that each period o f p time steps

is long compared to the settling time o f the system. Then the response during each

period can be characterized by the steady state response to the errors of the previous

period. Rewriting homogeneous equation (3.14) produces

z pE { z ) = [ \ - L { z ) G { z ) \ E { z ) (3.15)

The z pE ( z ) is a time shift o f the error signal p steps ahead. Equation (3.15) suggests

that w ithin the assumptions here, the frequency com ponents o f the error one period

ahead are related to the error components o f the current period according to equation

(3.11), now using j to denote the period rather than the repetition. Hence, for both

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
learning and repetitive control we wish to satisfy the same condition (3.12) for

purposes o f obtaining good learning transients. Again, this condition is a sufficient

condition for stability, but not a necessary one, as discussed in [6,18],

In the following sections we develop methods for use by the control system

designer to aid him in handling tradeoffs in satisfying the monotonic decay condition

(3.12) and obtaining good final error levels.

3.4 An Exam ple System for Num erical Investigations

In the following sections it will be helpful to have a specific example system

on which to apply the ideas discussed and observer the performance. We pick a

closed loop digital control system G(z) which is the discrete time version o f the

following third order system throughout this dissertation

f 2 \
( d 1 -) 6)0
U -t-c/J U " + 2£co0s + co; )

fed by a zero order hold with sampling at 100 Hz. The constants are d = 8.8

corresponding to a break frequency at 1.4 Hz, £ = 0.5. and con = 3 7 rad /sec

corresponding to an undamped natural frequency o f 5.9 Hz. This transfer function is a

good representation o f the closed loop response o f each joint o f a Robotics Research

Corporation robot [7,6], It is also sufficiently complicated to exhibit poor transient

behavior by many learning control laws and instability by many repetitive control

laws.

We pick a desired trajectory that is a summation o f many sinusoids going up

to the N yquist frequency:

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
y j ( k T ) = Y^an[\-zos((DnkT)~\\Q < k T <10 (3.17)
n=i

where the con are 0.2/z\2;r.4;z\6/r....J00/T . and the amplitudes are a„ = 80e_<u" . This

desired trajectory is shown in Fig. 3.5.

3.5 W hat Do W e Assume We Know About The System W hen We Design The

Learning or Repetitive Controller

An issue in learning and repetitive control is, how much information o f what

type do we assum e we know about the system so that a design can be made that will

have good learning transients? Here we say that it is reasonable to have the same

information that the control system designer has, since it may be the same person.

More specifically, we say that it is reasonable to assume that we have the following

information that we might simply obtain by running an existing feedback control

system and taking data to use in designing the learning or repetitive system:

I . We assume that we have a Bode plot o f the closed loop system behavior, i.e. we

know the steady state response of the control system to a sinusoidal input for all

frequencies up to Nyquist, both the amplitude change from input to output given by

|G(e'"r )|, and the phase change from input to output given by the angle m ade by

G(e'mT) with the positive real axis. Equivalently, we know G(e'a / ) for each

frequency up to Nyquist. If we obtain this information by applying a white noise

input, then our knowledge o f G(e'wT) for the higher frequencies may be somewhat

less certain. Even when direct measurements are made o f each frequency separately,

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
34

the uncertainty m ay be w orse at higher frequencies, because it may be hard to

generate a substantial am plitude to get above noise levels at such frequencies.

2. We can also have a m easurem ent o f the output error when w e applied the desired

trajectory as a command. Let U{z) = YJjz) in equation (3.1), corresponding to

applying the desired trajectory as the command. Then the error in the output

measured £>s(z) satisfies

EFB(z) = Ya( z ) - Y ( z )
= { \ - G ( <z))YJ { z ) - W { z ) (3-18)

The final expression converted to its frequency com ponents is one that we will need

in evaluating final error levels in one o f the options discussed below. And if we want

to know the frequency com ponents o f W(z) for a different option discussed below, we

can compute them because we know the frequency components o f YJ^z) and EFB{z).

and we have m easured values o f G(e,a,r).

Note that when there is a periodic disturbance W(z), the experimental

determination o f the frequency response needs to be done correctly. One can input a

signal such as white noise and record the data. Then input the same command

multiplied by some factor. Subtract the two inputs and the two outputs to eliminate

the repeating disturbances from the input-output relationship. One can then use this

differenced input-output data to create the frequency response.

3.6 The Design Approach

We have stated that we wish to satisfy the good learning transients condition

(3.12), and if w e do so for all frequencies, it will result in convergence to zero final

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
35

value o f the error. However, it is very difficult to satisfy (3.12) for all frequencies.

And even if we think that we have satisfied it. according to our model, it can easily

happen that there is extra phase lag at high frequencies that could not be determined

accurately. Following [18] the design process has two parts.

3.6.1 C om pensator Design

One first designs a compensator that attem pts to keep the plot o f the complex

numbers L(e'aT)G (e IQ,r) for co up to Nyquist frequency as much as possible within

the unit circle centered at +1 defined by (3.12). One m ight try to cancel the system

dynamics, but this is very often difficult. One can design compensators to influence

the phase. One very simple and effective method used in [36,12,14] is to use a linear

phase lead. The L(z) = (pz/ for y > 1 is a linear phase lead, and choice o f this simple

number can easily bring the plot o f L{emT)G(e,mT) inside the unit circle for nearly all

frequencies (see [22,39] for discussion o f automatic tuning methods for this

parameter). Figure 3.6 shows plots of <f>(emTY G ( e ”°T) for various values o f y . The

single learning gain <j) is equal to 1 in this figure. A lthough it is not visible to plotting

accuracy, the curves all go outside the unit circle. In the case o f y = 5 , the maximum

distance from +1 is 1.0051, going outside the unit circle by only 0.0051. This will

produce a long-term instability in repetitive control, and will produce a long-term

poor transient in learning control. Figure 3.7 shows the RMS tracking error versus

repetitions using this control law, decreasing by about 5 orders o f magnitude in

roughly 80 repetitions, and then it starts to diverge. In [18], several methods to handle

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
36

this persistent situation are discussed. These include: 1. use o f zero-phase low-pass

filtering. 2. stabilization due to quantization [13], 3. reidentification [8], and 4. use o f

matched basis functions [25,37,38]. Here we develop the zero-phase filtering option.

3.6.2 Zero-Phase Low-pass Filtering

Reference [6,7] introduces zero-phase low-pass filtering in learning control

and [12] does so in repetitive control. The objective is to cut out the frequencies that

go outside the unit circle from the learning process. By making the learning blind to

these frequencies we introduce a trade-off, we no longer try to learn all frequencies,

so we no longer aim at zero tracking error, and in exchange we produce good learning

transients in learning and repetitive control.

We can use a Butterworth low-pass filter to cut out the chosen frequencies. A

normal Butterworth filter will attenuate the frequencies above the cutoff, but will

introduce phase lags as well, and these disturb the plot sending it outside the unit

circle earlier. A zero-phase filter is created by filtering the signal o f interest forward

in time, then putting the output into the same filter but with the time sequence

reversed. The first application puts in phase lags, and the second puts in

compensating phase leads. There are some subtleties in extending the trajectories to

allow time for the filter transients to die away before reaching the trajectory segment

o f interest, both forward and backward.

We denote the zero-phase low-pass filter by F(z). Suppose that the original

Butterworth filter has transfer function Bw(z) with the amplitude Bode plot coming

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
37

from M F{e ",ir') -|£ » -(e,<ur) |. Then the frequency response o f the zero-phase version

o f the filter is just the amplitude change M F


2 (e,a,T) = F(e'“r ) , w ith zero change in

phase.

Suppose that we choose to filter the error to cut out the unw anted frequencies

before they go to the learning control law (3.7). The symbol Uj in equation (3.7) is an

accumulated signal associated w ith the sum o f all errors through repetition j - 1 (and

analogously for repetitive control for periods through period j - 1). I f the low-pass

filtering is not perfect, then some small amount of error in the unwanted frequency

range can get through. And this m eans that eventually this part o f the signal will grow

to dominate, and produce either instability or poor learning transients. Hence, we

must apply the filter to the accumulated signal to prevent such slow growth.

However, the accumulated signal contains the desired trajectory as the command

given in the Oth repetition. One may ask should this be filtered or not. We examine

both possibilities. In the first, we make a learning control law o f the form

UM (z) = F(z)[UJ (z) + L(z)EJ (z)] ,’ U 0(z) = Yd(z) (3.19)

In the second, we separate out the desired trajectory part o f the sum in (3.7) according

to

U j ( z ) = Yd (z) + U t j( z ) (3.20)

Ui (z) = F ( z )[U lj (z) + L{z)Ej (z)]: U 0 (z) = 0

and filter only the signal that is added to yJkT) by the learning process to form the

command each repetition.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
38

In repetitive control, the implementation o f zero-phase filtering is som ew hat

more com plicated. It requires that a time signal be filtered both forward and backw ard

in time, w hich is a noncausal computation. Here we suggest the following procedure.

Apply the repetitive control law version o f (3.7) for m repetitions. Then freeze the

repetitive control signal applying it each repetition until the zero-phase filtered result

becomes available. Starting with the last o f the m repetitions and continuing perhaps

through two repetitions after freezing the signal, we apply the zero-phase filter. By

applying the filter to three repetitions in a row with the same signal we can pick out

the results for the middle period, using the first and last for decay o f forward and

reverse filter transients. Once we have the filtered result corresponding to the right

hand side o f (3.19) or (3.20), we apply the result at the start o f the next period. Again,

we continue without filtering for m repetitions, and then repeat. In experiments on a

timing belt drive system reported in [12], the instability resulting from frequencies

going outside the unit circle, did not appear until around repetition or period 2650. In

this case, one m ight entertain applying the zero-phase low-pass filter every m = 2,000

periods.

The way in which the two options (3.19) and (3.20) appear in repetitive

control is rather subtle. W hen y^kT) is included in the accumulating signal as in

(3.19), we can use the block diagram as in Fig. 3.4. The block containing

L(z)f(z p - 1 ) is a difference equation needing initial conditions for the signal u(kT)

for time steps from k = I to p. We can use the actual command applied during the 0th

period w hich is the desired trajectory, and then the block produces the equivalent o f

(3.19). O n the other hand we could use zero initial conditions, in which case the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
39

accumulated signal does not contain y^JcT), and then we still have to add it on

afterward, as in Fig. 3.8.

3.6.3 Purpose o f this Chapter

In the above design process, w e are giving up zero tracking error to get

monotonic convergence — monotonic convergence to a non zero error level. One

asks, how do I design a learning or repetitive controller that produces monotonic

decay o f the error, and do so with as little final error as possible when the learning

process is complete. The answ er to this depends on the frequency ranges o f the errors

that one needs to eliminate. Here we develop methods that allow the designer to

predict the final error levels: 1. For each choice o f compensator, 2. For different

choices o f zero-phase low-pass filters, and filter orders, 3. For filtering the total

command signal as in (3.19) or filtering the command with the desired trajectory

removed as in (3.20), and 4. For different choices o f how often one does the zero-

phase filtering.

3.7 The Steady State E rror Level W hen Filtering The Total Command Signal

G oing Into The Feedback Controller

3.7.1 Iterative Learning Control

In order to predict error levels for the filtered learning control law (3.19), we

first find the difference equation for the error. Use equation (3.1) in the definition o f

error E/z) = Y^z) - Y{z) to obtain

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
40

Ej (z) — Yd (z) - G ( z) U j ( z ) - W ( z ) (3.21)

We now wish to com bine this with (3.19), and eliminate the U dependence so that we

have an equation for the error in terms o f the two inputs, Y^z) and W(z). To do this.

write the difference F /z ) = F(z)Ej.\(z) in order to create a term U j ( z ) —F (z)U ,_,(z)

which can be replaced by F( z)L(z)EJ_l (z) according to (3.19). The result is

Ej (z) - F ( r ) ( l - G (z)£(z))£,_, (z) = (1 - F(z))(Yd (z) - W{z)) (3.22)

The filter passes frequencies below the cutoff, essentially unchanged, so that

( l - F ( z ) ) is essentially zero in this frequency range. I f the filter were to make a

perfect cut off, at frequencies above the cutoff, F(z) is zero, and the forcing function

on the right is sim ply Yt/( z ) —JV(z). This remaining part o f the forcing function

produces a steady state particular solution. Examining the homogeneous equation, it

can be written as

£ ,( z ) = [F (z)(l -G (z )Z (z ))]F y_,(z) (3.23)

Hence, the monotonic decay condition becomes

|M j.(e,0,r )[1 - G(e,a,r)L{e,tuT)]| < 1 (3.24)

for all co up to N yquist frequency. It is clear that it is easy to pick a filter M } ( e ,a>r)

such that this inequality is satisfied for all co.

Once this decay condition is satisfied, we then wish to know what the final

error level will be. W hen this final steady state error level is reached, F ,(z ) = £ ;_,(z)

which we can call F vv( z ) . The transfer function from Yt/( z ) - W ( z ) to £„v(z) with z

replaced by e,a>r is

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
41

, % /2 \ r t
1 - M], (e'mT)[1 - G{e ioir )L{e,a“ )]
0 - 2 5 )

To find the steady state error level for each frequency com ponent, we decompose

Yi/(z) —W(z) into its frequency components and then m ultiply these input amplitudes

for each frequency by \$i (e “r )| for the corresponding frequency, to determine the

amplitude o f the steady state error at that frequency. W hen we are only concerned

with response to Ya{z) it is easy to decompose it into its frequency components. When

there is a repetitive disturbance PF(z), we are not likely to know what it is directly. In

this case we can determine its frequency com ponents as described below equation

(3.18).

3.7.1.1 A Simple Special Case

At first it would appear that a very sim ple design m ethod would simply use a

constant for the filter. As noted above, the values o f [1 —G(e ,a>T)L(e,a>T)] reaches a

maximum o f 1.0051 for the example o f (3.6) and (3.7), and using L(z) = <pz/ with

Y = 5 and a learning gain o f (p = 1. This suggests that all we need to do is put in an

F(z) = M p ( e “°T) = 1/1.0051, and we could produce a monotonically decaying error

for learning or repetitive control. Figure 3.9 shows the result. Due to the scaling in the

figure, one cannot see w here the curve starts, but the initial RMS error with feedback

only on repetition 0 is 2.5659. The RMS error at repetition 2 is 0.2605, and the final

value is 0.31445. This very small change from F(z) = 1 in Fig. 3.7 to 1/1.0051 in this

figure has made a very substantial degradation in the minimum value of the error

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
42

reached, but it did prevent the apparent divergence in Fig. 3.7. Figure 10 shows how­

s', (e'“r ) attenuates or am plifies the various frequency components in YJ^z) - W(z) in

producing the final value o f the error, for several different values o f constant

F ( r ) = M j,( e ,a,r) . We observe that this function is very sensitive to small changes in

the constant value, and can easily produce large am plifications o f errors in certain

frequency ranges. We conclude, that although the use o f a simple constant for the

filter F{z) at first appears attractive, that it is not likely to be effective at producing

low final error levels in practice. A more sophisticated filter is necessary.

3.7.1.2 Numerical Exam ple using a Butterworth Filter

We now use a 5th order Butterworth filter w ith a cutoff frequency o f 100

rad/sec (15.9 Hz). The filter cuts o ff before the system goes outside the unit circle in

Fig. 3.4. Since the filter is applied twice, the Mfr(e,a>r) attenuates above this cutoff

like a 10th order filter. Figure 3.11 shows the RMS error versus repetitions using this

learning law. As before the RMS error starts at 2.5659. This time it decays to a steady

error level o f 2.4684* 10-4 . The growth in the error in Fig. 3.7 has been eliminated,

with only a small increase in the steady error level above the minimum observed in

Fig. 3.7. Figure 3.12 shows how the choice o f the order o f the Butterworth filter and

the choice o f the learning gain influence the values o f j^i (e <ur) |, and this information

can be used to help pick these values in tuning the learning control law.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
43

3.7.2 Repetitive Control

N ow consider the use o f zero-phase filtering in repetitive control. As

discussed around equation (3.15). we use steady state frequency response to

characterize the response o f the system in the current repetition to the repetitive

control action taken in the previous repetition. This assumes that the settling time o f

the feedback controller is short compared to a period in the command/disturbance.

And this assum ption means that we can characterize the response in each period as if

it were a separate function from that in the previous period, and label the variables for

period j with the subscript j, just as we did in learning control. And this makes

repetitive control look very similar to learning control. The substantial difference is

that in repetitive control we cannot do the zero-phase filtering in real time, and use

the batch update procedure described below equation (3.20).

Suppose that we apply the repetitive control law L(z) for m repetitions, and

that there are r - m repetitions with the frozen signal before we apply the filtered

result and start the learning process again. Thus the process repeats every r

repetitions. Suppose we start with the error at repetition j = l r . Then the updates

each repetition obey

U ,r+I
(z) = U ,r (z) + L(z)E,r(z),...,Ufr+nt_ x (z) = U ,r+(„ _ 2 (z) + L(z )E(r+m_2 (z ) (3.26)

And then we wish to filter the next command signal, and it gets applied at repetition

j = V + l)r ,

(z) + ( 2 )]
U lM)r (z) = F(z){ U „ (z) + £ (z )£ ,„ , (z) + £ ( z ) £ ,„ , (z) + ••• + £ (z )£ „ „ „ , (z)] <3'27)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
44

During the first m repetitions, equation (3.15) applies which indicates that error at

repetition l r + a for a < m is related to the error at repetition ir by

E ,r + a (-) = (1 - G ( z )L{ z ) Y E,r(z) (3.28)

Substituting into (3.27) gives

£ W (*) ~ E { z ) U tr (z) = F(z)L(z)[ 1 + (1 - G(z)L(z)) + - + (1 - G ( z ) L ( z ) ) ] E (r (z)

" l - d - G ( z ) L ( z ) ) lr+m'
= E { z ) L ( z) E,r{z)
G (z)L(z)
(3.29)

N ow consider the difference £ (,+1)r(z) —F(z)E,r(z) using Ej( z) = Yd - G U j —W . so

that the left hand side o f (3.19) appears. Then substitute the right hand side for this

term in order to eliminate Uj dependence. The result is

E{e+l)r(z) - F(z)( 1 - G { z ) L { z ) T Efr( z ) = (1 - F{z))[Yd{ z ) - W { z ) } (3.30)

This equation is analogous to equation (3.22). This time the frequency components

outside the unit circle are allowed to grow for m repetitions, and the filter F(z) must

be picked to attenuate this accumulated growth to a number less than one. The

equation analogous to (3.24) is

\Ml.{e'°>T) [ \ - G ( e ' a’r )L(e'“T) r \ < 1 (3.31)

and satisfaction o f this produces monotonic decay o f the frequency components o f the

error for the errors after each filtering, at &*,(£ + l)r,(£ + 2 ) r ,.... The frequency

transfer function analogous to G.25) is

£ (e™) = I-M W *)
2 1 —Mj-(e'“7 )[1 —G{e,iaI )£(e""7)]"'

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
45

M ultiply the m agnitude o f this for any frequency tim es the am plitude o f the

com ponent o f Yd (z) —W (z) at that frequency, and you get the am plitude o f the steady

state error at that frequency.

3.7.3 Periodic Filtering in Learning Control

In the learning control discussion above, we applied the zero-phase low-pass

filter every repetition. Experiments show that the error level m ay stay low for many

repetitions, and one w ould not feel the need to apply the filter every repetition.

Instead one m ight apply it every 100 or 1000 repetitions, as appropriate, to keep the

RMS error from starting to grow. In order to predict the steady state error levels for

this process for any choice o f the number of repetitions between applications, the

same com putations as in the repetitive control case above apply.

3.8 The Steady State Error Level When The Desired Trajectory Part o f The

Com m and Is Not Filtered

3.8.1 Iterative Learning Control

In the previous section we filtered the entire comm and signal that goes to the

feedback control system, as in equation (3.19). In this section we consider the

differences when we use equation (3.20) instead, taking out the desired trajectory

from the signal that is filtered. In place o f (3.21) we have

EJ (z) = Yd (z) - G( z) UJ (z) - W ( z )


= (1 - G{z))Yd (z) - G { z ) U (z) - W(z) (3'3 J)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
46

Following the sam e procedure as before, this time to elim inate U L dependence, we

use (33) in EJ{z) —F { z ) E /_l{z) in order to create a term U Lj (z) —F( z ) U Lj_x{z) and

replace it by F ( z ) 0 ( r ) £ j_ ,( r ) . The result analogous to (3.22) is

£ j (z) - F{z){ 1 - G (z)L(z))£,_, ( z ) = (1 - F (z))[(l - G(z))Yd (z) - W(z)} (3.34)

Examining this equation, the discussion o f monotonic decay is the same as in

equation (3.23) and (3.24), but the steady state error addressed in (3.25) has som e

new aspects, discussed in the following points:

1. Suppose that we have made an experimental run giving the desired trajectory' as the

com m and to the feedback controller, and measured the error in the output. Then

according to equation (3.18), we have a direct m easurem ent o f EFB(z) which is

m ultiplied by 1 - £(z) to form the forcing function in (3.34). Then we can find the

amplitudes o f the frequency components in E,,B( z ) , and multiply by the magnitude o f

Sx(e,a>T) in equation (3.25) for that frequency component, and obtain the amplitude o f

the steady state error for that frequency. Thus, the frequency transfer function

S ,(e'“r ) applies directly to the measured error using feedback only, whereas in the

previous case we had to use the measured error and com pute W(z) in order to know

the input Y\,{z) —W ( z ) . Thus we use the same S l(e"u l) in both cases, but we apply it

to different signals, and hence we will obtain different steady state errors.

2. Suppose our main interest is the ability o f the system to eliminate the effects o f

disturbances W(z), and perhaps YJz) = 0, as often happens in repetitive control. Then,

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
47

equation (3.34) and (3.25) become indistinguishable, and it does not m atter which

approach w e use. The results will be the same.

3. Suppose that we do not expect any disturbances that repeat each repetition or

period, and our concern is to know the steady state error in following the desired

trajectory. Then we find the amplitudes o f the frequency components o f Yd(z). and

this time m ultiply by the magnitude o f the following frequency transfer function for

the same frequencies, in order to know the steady state error for each frequency

cw -'T x [ l - M ; - ( e ^ ) ] [ l —G { e - T)}
\ - M l . { e ,0lT) \ \ - G { e ,0,T)L(e‘mT)} ’

Figure 3.13 shows the result o f using this learning control approach o f equation (3.20)

on our example problem. The RMS error stabilizes at 3.519x10"' which is

substantially better than the steady state error level for the approach o f (3.19). It is

very close to the minimum error level observed in Fig. 3.7. So the approach here is

able to reach essentially the same minimum level as without the filtering, and then

stop the growth thereafter, maintaining this error level indefinitely.

3.8.2 Comparison With the Previous Case

One now wants to know which approach is better, to filter the signal including

the desired trajectory as in (3.19) or to remove it as in (3.20). In the above example, it

was the second approach. But the answer depends on the situation. If the point o f the

learning or repetitive control process is to eliminate repetitive disturbances, then it

makes no difference which one uses. If the point is to eliminate the deterministic

errors o f the feedback controller in following a tracking command, i.e. the error in the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
48

particular solution, then which one is better depends on the dom inant frequency range

o f in the command. Figure 3.14 shows the ra tio

(3.36)

If the dom inant frequency content o f the com m and is in a range for which this plot is

above unity, then it is best to filter the desired trajectory as in (3.19). If the dom inant

error is a frequency range where this plot is below one, then separate the desired

trajectory from the filtering process as in (3.20).

3.8.3 Repetitive Control

N ow consider the repetitive control version o f the filtering with the desired

trajectory removed. We can parallel the developm ent o f equations (3.26) through

(3.32). The only change is that again we ad ju st £ y(z) according to equation (3.33).

Examining the development we see that th e only change is to replace Y^z) by

(1 - G ( z ) ) Y j ( z ) (and replace £\(z) by E [ f {z). but these get eliminated). Then the

equations analogous to (3.30) and (3.32) becom e

E„_ 1)r(z) - E ( z ) (l- G ( z )£ ( z )) “ Efr(z) = (1 - E (z))[(l - G ( z ) ) ^ ( z ) - W(z) ]


n k/ f - ( a ,oir vi n _ n ( „ ,o>r m (3.37)

Again these equations also apply when u stn g periodic filtering in learning control

rather than filtering every repetition.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
49

3.8 Conclusions

In this chapter we have developed analytical methods that guide the designer

o f learning o r repetitive controllers in getting m onotonic decay o f the tracking error in

repetitive situations, and getting the best final value o f the error when the learning

process has converged to steady state. The frequency transfer functions in (3.25).

(3.32), and (3.35-3.37) predict the steady state error levels for different learning and

repetitive control approaches. These can be used to adjust the compensator used, such

as making the choice o f the linear phase lead y , and to pick the learning law to use or

to adjust a learning gain such as $ . And these equations allow the designer to make

the best choice o f whether to include the desired trajectory in the signal to be filtered.

In the case o f repetitive control the equations are used to decide how often one must

apply a filter in order to keep the RMS error from growing. The approach developed

here is practical in the sense that all the learning or control system designer needs to

know is experim entally measured closed loop frequency response information for an

existing feedback controller, and the error it produces in following the desired

trajectory w hen it is given as the command. From this he can design a learning or

repetitive controller that will substantially improve the tracking performance over that

o f the feedback control system, by iteratively adjusting the command given to the

feedback controller.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
50

Vds)

Ftz)
D/A
A/D
ZO H

A/D

Fig. 3.1 Typical digital feedback control system.

viz)

♦O

W{z)

Y(z)
*o

Fig. 3.2 Converting disturbance Vc(s) to a discrete time output disturbance.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
51

60

40
RMS tracking error

to-'

•20

,-40

,-60
0.5 2.5
Repetition number

Fig. 3.3 The learning transients during the convergence to zero tracking error in
integral control based learning.

mz)

YAz) + E (z) U(z)


L (z)
G(z)

Fig. 3.4 Repetitive control applied to the closed loop control system G(z).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
52

100

o
oo 60
C3

40

20

Time(sec.)

Fig. 3.5 The desired trajectory for the numerical simulations of learning control.

0.8

0.6

0.4

G
.
G3 0.2

0.2

-0.4

- 0.6

- 0 .;

-0.5 0.5
real part

Fig. 3.6 Nyquist plot of product of the learning law and the closed loop transfer
function, (fc rG(z), for y = 1.2. 3. 4. 5. and 6 .

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
53

o
C3

i o '5
0 100 200 300 400 500
Repetition number

Fig. 3.7 RMS tracking error versus repetitions using a learning controller with a
linear phase lead of y = 5 .

VJ.Z) W(Z)

YAz) + E(z)
Uz)
G(z)
4I zT-l

Fig. 3.8 Repetitive control with the desired trajectory removed from the
accumulated signal by setting the repetitive controller initial conditions to zero.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
54

5o 1CT
C3

in
2
o'

to'M_ _ _ _ _ _ _ _ i_ _ _ _ _ _ _ _ i_ _ _ _ _ _ _ _ c_ _ _ _ _ _ _ _ ,_ _ _ _ _ _ _ _
0 100 200 300 400 500
Repetition number

Fig. 3.9 Learning control using the simplest zero phase filter, constant F(z) =
1/1.0051.

F(r) =1/1.0051
1. 6 - ... F(z) =1/1.01
— F{z) =1/1.015
.4-
1.2

0.8
0. 6 -
0.4-
0.2

40
Frequency (Hz)

Fig. 3.10 The values o f l-SiCe"^)! versus co for different choices of F(z). F{z) =
1/1.0051, I/1.0I,and 1/1.015.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
0 100 200 300 400 500
Repetition number

Fig. 3.11 RMS tracking error versus repetitions when the entire command each
repetition goes through the zero-phase low pass filter.

0.8

04

0 .4 -

0.2

40
Frequency (Hz)

Fig. 3.12 The values o f )| versus co when using the n th order Butterworth
zero-phase filter on the entire command with different learning gains (f>. Curves,
left to right. («, $ : (5, 0.25), (5. 1). (10. 0.25), (10, 1).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
56

I0'5t________ .________ i________ «________ ,________


0 100 200 300 400 500
Repetition number

Fig. 3.13 RMS tracking error versus repetitions when the desired trajectory part of
the command is not filtered.

0.8

0.6

0.4

0.2

Frequency (Hz)

Fig. 3.14 Comparison o f including or not including the desired trajectory in the
filtered signal.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
57

C hapter 4

The W aterbed Effect and Iterative Learning C ontrol U sing Zero-

Phase Filtering

4.1 Introduction

Iterative learning control (ILC) seeks to eliminate determ inistic errors in

repeatedly following a specific desired trajectory, by learning from previous

experience. The original motivation was for eliminating errors o f robots in executing

the same trajectory many times. There are two classes o f errors. One is the fact that

for nearly all tracking commands, the response o f a feedback controller is not equal to

the command. The second is that in many situations, there are repeating disturbances

every time the same command is executed, for example, the gravity torque on a robot

link as it moves along a trajectory. Iterative learning control applies to situations in

which the system, e.g. the robot, returns to the same starting point before each

execution o f the command. Looking at the error in following the trajectory in the

previous repetition, the command is adjusted in the next repetition, in order to

decrease the tracking error, and it is possible to have learning control converge to

zero tracking error.

In classical or digital feedback control theory, there is a fundamental

result, the Bode integral theorem [30], which can be interpreted as the waterbed

effect. The waterbed effect was first recognized by Francis and Zames [9] as they

studied the feedback constraints imposed by open right h alf plane zeros o f the plant

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
58

in the context o f H-infinity optimization. Looking at the frequency response of a

typical feedback control system, for the transfer function from com m and to resulting

error, the waterbed effect says that if the error is attenuated substantially by the

feedback control law over some frequency range, then it w ill be amplified

substantially over some other frequency range. In learning control we seek to have

zero tracking error at all frequencies, although to make design easy and

straightforward, and to produce robustness o f the algorithm, it is usually advisable to

cut off the learning process above some frequency. In this paper w e investigate in

what way the Bode integral theorem type o f thinking applies to large classes o f

learning control laws. We investigate how ILC laws are able to bypass the waterbed

effect, to potentially produce zero error at all frequencies, or to substantially improve

error over a range o f frequencies without paying for the improvement by amplifying

errors in some other frequency range. This includes introducing more zeros than

poles, modifying the gain in the sensitivity transfer function, and introducing poles

outside the unit circle with noncausal filtering.

The waterbed effect is a steady state frequency response result. Iterative

learning control is concerned with executing a finite time desired trajectory. Hence,

the results we obtain here will relate to the execution of the desired trajectory for that

portion that is independent o f initial conditions, i.e. for that part o f the trajectory after

a number o f time constants have elapsed. This means that we need not be concerned

with initial conditions, and transfer function modeling is sufficient.

Consider a digital feedback control system with closed loop transfer function

G (z ). Then the response to input U / (z) is written as

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
59

YJ( z ) = G ( z ) U J(z) + W(z) (4.1)

w here tV(z) is any repeating disturbance that appears every tim e we execute the

trajectory. Subscript j indicates the repetition number. The desired trajectory is Yd (z)

and define the tracking error as Ey(z) = Yd (z) - Y ( z ) . In this chapter we follow ILC

laws from chapter 3

UJ^ z ) = UJ{z) + ^ E J{z) (4.2)

Write (4.1) for j +1 and for J, take the difference, and apply the learning law

(4.2), to obtain the error update equation from one repetition to the next

£ y+,(z) = [l-< fe 'G (z )]£ ,(z ) (4.3)

The square bracket contains a transfer function from the error at one repetition to the

error at the next repetition. Substituting z = exp(icoT) (T is the sam ple time) produces

the frequency transfer function, whose magnitude at any frequency determines the

attenuation or am plification o f the amplitude o f the error at that frequency from one

repetition to the next. If

(4.4)

for all frequencies up to Nyquist. then all amplitudes o f the error decay monotonically

with repetitions, for the part o f the trajectory governed by steady state response. In

[19,6] it is show n that this is a sufficient condition for stability o f the learning

process, and convergence to zero tracking error for all time steps o f the trajectory.

Because o f its monotonic decay property for the steady state part o f the trajectory, it

is suggested in [6] as a good learning transients condition.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
60

4.2 The Standard W aterbed Result

Consider a typical digital feedback control system with unity feedback. After

converting the plant Laplace transfer function fed by a zero order hold to a z-transfer

function, and com bining it with the controller transfer function to form Go[{z) in the

forward part o f the loop, the transfer function from command to error is

S(z) = 1/(1 + Gol( z ) ) . This transfer function has two special properties: the degree o f

the num erator polynomial and the denom inator polynomial are the same (call it n),

and the lim it as z goes to infinity in any direction is one.

Theorem 1 (Bode integral theorem): Let S(z) be a transfer function having the

above two properties. Suppose also that it corresponds to an asymptotically stable

system. Define ln*|/?| = max(0,ln|/?j) for any com plex num ber J3, and let the zeros of

S(z) be z,. for i = 1,2,3,— - Then

(4.5)

A proof is given in [30] making use o f the Poisson integral that requires

stability. Below we generate a proof o f a m ore general theorem that has this result as

a special case, and does not require stability.

In (4.5) one can let go = 0 / T to represent the radian frequency. Hence, (4.5)

says that the logarithm o f the amplitude Bode plot o f the sensitivity transfer function

S(z) , averaged over all frequencies up to N yquist is given by the right hand side. If

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
61

there are no zeros outside the unit circle, then the right hand side is zero. This

produces the waterbed effect. If one decreases the error substantially in a frequency

range, then there m ust be some other frequency range for which the error is amplified.

The result is in logarithm space. To illustrate the implication, suppose that the error

up to one h alf Nyquist frequency were decreased by a factor o f 10, then the error

above this frequency m ust be increased by a factor o f 10. It is not the error itself that

averages to zero, it is the factor multiplying the error that averages to zero.

Perhaps the majority o f digital control system s will have zeros outside the unit

circle. W hen one creates the discrete equivalent o f a plant fed by a zero order hold,

and there are at least two more zeros than poles in continuous time, then there will be

at least one zero outside the unit circle if one samples fast enough. The presence o f

such a zero in the Bode integral simply increases the right hand side to some positive

number. Hence, zeros outside the unit circle m akes matters worse, there will be more

amplification than attenuation in the average.

In any case, there is no way in which one can get a negative number on the right

hand side. This means that typical feedback controllers improve performance by

producing significant attenuation in a frequency range where the important error

occurs, and amplifying in parts o f the frequency spectrum where there is very little

signal to start with, so that the amplification can be tolerated.

It is the purpose o f this paper to investigate in what ways the Bode integral

theorem type o f result applies to iterative learning control. More specifically, we seek

to understand how it is possible for linear ILC laws to produce large attenuation of

error over most or all o f the frequency range up to Nyquist, without being subject to a

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
62

waterbed effect, i.e. without having to pay for this benefit by amplifying errors

somewhere else.

4.3 A Robot Example

In order to illustrate the developments o f the follow ing sections, a third order

model o f the input/output relationship for each jo in t o f a Robotics Research

Corporation robot is used [6,7]. The continuous tim e m odel is given in equation

(3.16).

Note that one o f the transfer functions that w e w ould like to consider with

respect to a Bode integral theorem type of result, is the transfer function in (4.3).

Suppose that (f>= 1 and y = 1 (and G(z) has one more pole than zero). This

corresponds to the most common learning control law based on pure integral control

action applied at every time step in the discrete repetition domain. The transfer

function in (4.3) then has the same number o f poles as zeros, but there is a coefficient

in front depending on the learning gain. The presence o f this coefficient means that

the conditions o f the Bode integral theorem are not satisfied. Consider a / > 1. This

produces more zeros than poles, and again the conditions o f the theorem are not

satisfied.

4.4 Generalized Bode Integral Theorem

In order to examine the frequency response characteristics of ILC. it is

necessary to create a much more general version o f the Bode integral theorem.

Consider a general transfer function

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
63

R(z) = K r 7- —~ ~ 7 "—~ ~ (4-6)

This time we allow the degree o f the numerator polynom ial to be different

than the degree o f the denom inator polynomial, n is not necessarily equal to m. And

in fact we will allow more zeros than poles. Also, any non zero gain K r is allowed, so

that even if m was equal to n, the limit as z goes to infinity is not necessarily one. And

finally, we do not require that all poles be inside the unit circle. We create the

analogous result to the Bode integral theorem, for this general situation.

Theorem 2 (Generalized B ode integral theorem, GBI): For the R(z) o f equation

(4.6)

1 ~ m n
- J l n | « ( e " ) |e / ^ = l n | ^ | + £ l n * | 2/| - £ l n * | / » l| (4.7)
^ 0 »=i '=1

As before, zeros outside the unit circlemake this logarithmic average o f the

frequency response larger,corresponding to more error in m ost uses o f the theorem.

On the other hand, having a K r that is less than unity in magnitude decreases the

value o f the integral, and having poles outside the unit circle decreases the values o f

the integral. In the following sections we will see how these two effects are used by

learning control laws to get improved performance over a broad band o f frequencies.

Proof: Because o f the possibility o f poles outside the unit circle, we need to

use a different approach to obtaining a proof, making use o f a mean value theorem in

complex analysis. First rewrite the integrand in (4.7) as

ln\R(z)\ = ln|/f„ | + In|z - r, | + • ••+ ln|r - z,„ \ - ln|z - p x| --------ln|z - p H| (4.8)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
64

Substitute z = exp(/<9) and integrate each term from —jt to JrTC, and divide by 2 k .

Because j/?(e'°)| and |/?(e~'s )|are equal, the stated integral o f the left hand side of

(4.8) becom es the left hand side o f (4.7). The integration o f the ln|^fR| term leaves it

unchanged, equal to the first term on the right o f (4.7). It remains to demonstrate that

the integral o f the first order terms produces

.T

(4.9)

m aking use o f concepts in [17]. This dem onstration is divided into three cases, when

P is inside the unit circle, on the unit circle, or outside the unit circle.

C ase 1: Let p = 6 e x p ( /» with 0 < b < 1. Consider the integral

(4.10)

where CQ is the unit circle about the origin. Note that In(l + z) = f - i f 2+ i f 3 — - is

convergent for |i| < 1 (and convergent on the unit circle except at f = - 1). which

implies that the integrand is analytic in the disk o f radius \z\ < 1/ b larger than the unit

circle. Hence, the integral in (4.10) is equal to zero. Using z = exp {id) on the unit

circle, and converting the variable o f integration from 6 to a new 9 equal to 9 +cp.

one can write (4.10) as

~ jt Z
0 = / Jln(l - be'6 )de = i |R e[ln(l - be'0 )]c/ 0 - Jlm [ln(l - be'0 )]d9 (4 . 11)

The first integral on the right is the imaginary part o f (4.10) which must be zero.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
65

N ote that it is closely related to the integral of interest, because

Injl —be'01= Re[ln(z —be'0)]. It is rewritten as

(4.12)

w hich indicates that its integral is the negative o f the integral o f interest. But the

integral is equal to zero which completes the proof for this case.

C ase 2: When /? = exp(/<p), w e cannot apply the same technique as above

since the logarithm series does not converge. In place of (4.10) consider the integral

(4.13)
(

around the contour in Fig. 4.1. To form this integral we have absorbed the phase (p

into 6 as in the previous case. Again, we know this is equal to zero. To evaluate the

first integral, substitute z = I + se‘° and dz = iee'°dO . Taking the limit as s —> 0 . and

noting that s l n s - > 0, again produces the result that the integral (4.9) is zero for this

case.

Case 3: When J3=bexp(i(p) with b > 1. ln(z —J3) is analytic in a disk

jzj < 1+ e for some e > 0 . The real and imaginary parts o f an analytic function are

harmonic functions, i.e. ln |z -/? | is harmonic in this disk. The mean value theorem

for harmonic function f ( z ) on the unit disk states that

1
(4.14)

which produces (4.9) using f ( e ' ° ) = lnje '0 - /?[.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
66

4.5 Finite Tim e Trajectories and T he Frequencies in The G BI Theorem

N ote that the generalized B ode integral theorem (as well as the Bode integral

theorem ) presents a logarithmic average over a continuum o f frequencies from zero

up to the Nyquist frequency. W hen applying digital learning control, the desired

trajectory has a finite number o f tim e steps, and this implies that the error record for a

repetition contains information only about a finite number o f frequencies, i.e. the

frequency associated with the total time o f the trajectory as period, and all harmonics

up to Nyquist. Hence, in practice, the only meaningful frequencies are a finite set of

discrete frequencies. As the num ber o f time steps in the trajectory is increased, the

num ber o f discrete frequencies up to N yquist that are contained in the sam pled data

becomes larger, tending to a countable infinity. The mathematics o f the developm ent

o f the integral theorem does not have any information about the length o f the finite

time trajectory o f interest, and produces an average over a continuum o f frequencies.

Technically, the theorem average only applies as the trajectory length tends to

infinity, so that all frequencies are present. We must expect such a lim itation, since

the theorem is concerned with steady state frequency response, and technically no

system reaches steady state except in the limit as the trajectory time tends to infinity.

4.6 Implications for Causal ILC Laws

One large class o f learning control laws use a causal learning law. By this we

mean that the degree o f the num erator in (pzrG { z ) is less than or equal to the degree

o f the denominator. This implies that the learning law does not make use o f error

terms in the previous repetition that are in the future to the current time in the present

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
67

repetition, but adjusted forward by the usual one time step delay o f the system. Causal

learning controllers are formed by using a causal filter or com pensator with the

num ber o f poles greater than or equal to the number o f zeros, and by picking the

value o f y less than or equal to the sum o f delays through the feedback system and

compensator.

Consider the standard integral control based learning applied to the robot

example, using a learning gain o f (p = 1 and y = 1 to look at the error one step ahead

in the previous repetition. Computing the transfer function o f (4.3) gives 3 zeros and

3 poles, all within the unit circle, and it has gain K r for equation (4.6) equal to

1 - <pKa , where K G is the corresponding Kr value for the discrete time zero order

hold version o f (3.16). The result is that the integral o f the GBI theorem (4.5) is equal

to ln|/CR| = -0 .00178. Although the standard Bode integral theorem does not apply,

which would have made the integral zero, the logarithmic average represented by the

integral is still only slightly negative. And it is not sufficiently negative to keep from

amplifying certain frequency components in the steady state frequency response.

Although the learning law is asymptotically stable, this am plification makes very

poor transients during the learning process [19,21]. Note that the value o f K c is

0.0018. Thus, a very large learning gain would be required to significantly reduce the

integral and reduce the waterbed effect. Going significantly above a learning gain o f

unity is not advisable. A learning gain o f unity totally corrects the DC error in the

next repetition, since the DC gain o f (3.16) is unity. A learning gain between 1 and 2

would still produce convergence o f the DC error, but with an alternating sign each

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
68

repetition. And a learning gain greater than 2 would cause the DC com ponent o f the

error to grow each repetition. This statement, like the generalized Bode integral

result, applies to the steady state part o f the trajectory. It implies bad transients during

learning, but it does not imply instability [19], a gain approaching 100 is allowed

before there is mathematical instability. We see that although this learning control

law is not subject to the usual Bode integral theorem constraint, the usual waterbed

effect, it is still not a practical result in most applications.

4.7 GBI Theorem and Noncausal ILC

Learning control is fortunate that the data processed for the update o f the

control action comes from the previous repetition. The time step o f m ost importance

in the previous repetition is that one corresponding to the current tim e step, but

shifted forward one step. But it is possible to make the learning law depend on steps

future to this in the past repetition. Reference [36] discussed the fact that this allows

one to use noncausal digital signal processing techniques, including using the linear

phase lead y . Such learning laws have the property that (p zf has more zeros than

poles, with the excess being larger than one (when G(z) has the usual property of

one less zero than pole). In place o f the 1 — < pK G o f the previous section for the K r of

equation (4.6), we get — <f>Ka . N ote that this can be made arbitrarily small if we are

willing to learn sufficiently slowly. This gives one considerably more control over the

logarithmic average o f the frequency response o f equation (4.7). M aking this product

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
69

o f gains sufficiently small can make the logarithmic average go toward negative

infinity, or the frequency response amplitude go toward zero for all frequencies.

As an example o f noncausal learning control, consider the ILC o f equation

(4.2) with a learning gain 0 = 1. a phase lead o f y = 5 (including one step delay o f the

system), and 0 = 1, applied to the robot problem. Then 1—<pzr G{z) has three poles

and seven zeros. Four o f these zeros are outside the unit circle. The logarithmic

average o f equation (4.7) is —0.05998. Again, the usual waterbed result does not apply

to this learning control law, and we can have a negative logarithmic average.

One might want this average to be more negative. The learning gain used was

0 = I and this produced a l n |^ R| = ln|0£f;| term in (4.7) equal to -6.3296. Decreasing

the gain to 0 = 0.1 makes this term equal -8.6322, and the logarithmic average o f

(4.7) is decreased to -0.80221. This is a substantially improved attenuation o f the

error every repetition. Further decrease o f the learning gain to K R =0.01 makes

ln|/ffl | = —10.9348, but the logarithmic average o f (4.7) becomes worse, equaling -

0.0005222. To understand the observed behavior, note that the learning gain is part o f

the coefficient o f the highest power in the zeros polynomial. As this coefficient tends

toward zero, at least one root tends to infinity. Therefore, decreasing the learning gain

improves the first term on the right o f (4.7), but it can make the second term worse.

Hence, from the point o f view o f the logarithmic average, there is an optimum value

for the learning gain in the robot example.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
70

4.8 Low-Pass Filter Learning C utoff

Experience indicates that it is not easy to satisfy the good transients condition

(4.4) all the way to Nyquist frequency. There will normally be some high frequency

interval in which the magnitude o f the transfer function (4.3.4.4) exceeds unity. This

can cause a slow or fast long-term growth o f the error before convergence occurs

[19].

As a result, low-pass filtering is introduced in [6.7] to cut o ff any frequency

range that violates (4.4) (see also, [19,29,21]). This is also a technique for producing

robustness o f ILC laws. Since ILC normally aims for zero tracking error, the

m odeling errors at high frequencies, no matter how small are not ignored, and they

can be extrem ely important to the convergence properties. This fact is in contrast to

normal feedback control, where m odel errors are unimportant when they are at high

frequencies w ith small output, where the frequencies are far above the zero dB

crossing line where stability is determ ined. Modeling errors at high frequency are

normally associated with parasitic poles, or unm odeled modes, and m athematically

can be described as singular perturbations. The filtering can be applied to cut out

these frequencies from the learning process.

There are various types o f filtering. Here we consider IIR filters, and in

particular, Butterworth filters. To im plem ent such a filter, one applies a chosen order

Butterworth filter to the data set for one repetition, and then applies the same filter to

the resulting output sequence, but do so running through the sequence backwards in

time. Each application produces attenuation above the cutoff, and the reversed time

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
71

application introduces a phase lead that cancels the phase lag o f the original

application.

The transfer function for such a filter is given in [3] as

(4.15)
V'
1+
l + z -I

where n is the order o f the original filter, c -cot{cocTI T) , and coc is the filter cutoff

frequency. For exam ple, when n is 3 the above equation is produced from

1
H{ z) = (4.16)
1+ c
1—z

l + z
_-i -t 'N

As a product o f H(z~ ) and H ( z ) , the filter clearly has as a frequency transfer

function with z - exp(icoT) substituted.

Reference [21] gives two options for application o f the filtering. In the first,

the control signal in (4.2) is filtered before application, producing the learning control

law

U J+l (z) = F( z)\ Uj (z) + K l«K z ) z ' E j (z)] (4.17)

with U0( z) = Yj ( z ) . In the second case, the desired trajectory is separated from the

learning control adjustm ents, and only the learning control adjustm ents are filtered

U lj +x(*) = n z ) [ U Uj{z) + K [ ® ( z ) z ' E J (z)]


(4.18)
u l.q = 0; u j M( z ) = yj ( z ) + u Uj +1( z )

The choice o f w hich option can be optimized based on the error amplitude frequency

distribution as described in chapter 3 and [21,19].

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
72

4.9 GBI Theorem and Zero-Phase Filtering

H aving introduced zero-phase filtering, we are no longer asking for zero error

in the limit. Hence, in addition to the transfer function o f the error from one repetition

to the next, it is o f interest to look at the transfer functions from the command to

resulting error, and from the repeating disturbance to error, to see what is implied by

the logarithmic average (4.7). The robot examples below use a 5th order Butterworth

filter, a learning gain o f 1, and y = 5.

Consider first the zero-phase filtering as in (4.17). Following [21], combine

the definition o f error, equations (4.1) and (4.17), in such a way as to eliminate t/y (z) .

This produces

EJ (z) - R, (z )£ y_, (z) = [1 - F(z)\[Yd (z) - fF(z)] (4.19)

Paralleling the thinking in [6], we can sum the geometric series involved to obtain

EJ(z) = [Rl ( z )]7 E0( z ) + 5, (z)[Trf (z) - W{z)]


RL{z) = F(z)[\ - K [ <t>(z)zrG(z)] (4.20)
5 ,(z) = [ l - F ( z ) ] / [ l - ^ ( z ) ]

To design a learning controller, one adjusts the cutoff frequency o f the zero-phase

filter to make [/?, ( e ,a)‘ )| < 1 for all frequencies up to Nyquist. Note that H(z) is

analytic within a disk about the origin, which contains the unit disk. And therefore

H(z~l ) isanalytic outside some disk with radius less than unity.This makes the

product analytic in an annulus containing the unit circle. Figure 4.2 plots RL(z) fo rz

going around circles o f different radii, and shows that it is less than one in magnitude

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
73

for radii in the range 0.9701 < |z| < 1.1499. Once convergence is obtained in such an

annulus, then its definition can be extended by analytic continuation.

In the limit as the repetitions go to infinity, the first term on the right will go

to zero for all z in the annulus. Then Sjfz) represents the transfer function from

command to the resulting error after learning is complete, and it also is the transfer

function from any repeating disturbance to the resulting error. ■

W hen the zero-phase filtering is applied as in (4.18) the equation

corresponding to (4.19) becomes

EJ (z) - Rl ( z ) ( z ) = [1 - F (z)][( 1 - G(z))Yd (z) - W{z)} ( 4 . 2 1)

and there is a modified sensitivity transfer function from com m and to error

5 3( z ) = [ 1 - G ( z ) ] 5 , ( z ) (4 .2 2 )

while the transfer function from disturbance to error remains the same. Figure 3 .1 4 in

chapter 3 shows what the m ultiplying factor l - G ( z ) looks like as a function o f

frequency up to Nyquist. Depending on where the majority o f the error is in the

frequency spectrum, one o f the two approaches will be best for the final error level.

The logarithmic average for the two sensitivity functions S j(z) and S~(z) are

identical. To see this, note that the log o f a product is the sum o f the logs, that the two

transfer functions differ by the factor l - G ( z ) , and that this factor satisfies the

conditions o f the standard Bode integral theorem. The logarithmic average for both o f

them is —2.88177, which represents a substantial improvement over any feedback

control method that is subject to the standard Bode integral theorem. Although both

integrals give the same logarithmic average, the error is distributed differently with

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
74

frequency in accordance with Fig. 3.14. Figure 4.3 gives a plot for S', (r) as a function

o f frequency. The plot o f S3(z) is too sim ilar to show. The plot shows attenuation to

'‘zero” below the cutoff, and “no” am plification above the cutoff -- the perform ance

one aim s to achieve. As one might expect, examining the numbers from the com puter

shows that the part o f the curve above the cutoff is actually slightly larger than one.

N ow consider how this perform ance is accomplished. For S,(z) there are 13

zeros and 17 poles. O f the zeros, three are clearly inside the unit circle and the

rem aining 10 numerically differ from one, either larger or smaller, only in the 9th or

10th decimal place. So there is no real adverse effects from the zeros term in (4.7).

The new aspect o f the filtering approach is that there are now poles outside the unit

circle. Since the system is not causal, this does not mean instability. O f the 17 poles, 9

are outside the unit circle. As a result, the poles term in equation (4.7) contributes -

15.2572. On the other hand, the gain term is given by i n | = 12.3754, making the

overall logarithmic average (4.7) equal to —2.88177.

4.10 Conclusions

In this chapter we have shown that iterative learning control law's are able to

bypass the waterbed effect that applies to nearly all feedback control systems. This is

accomplished by two techniques. One is to arrange that the gain K r be something

other than unity for the appropriate sensitivity transfer function, from com m and or

disturbance to error, or from error at one repetition to error at the next. The second is

that ILC can use noncausal filtering, and this allows one to have poles outside the unit

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
75

circle without producing instability. A generalized Bode integral theorem is presented

here to handle the new situations introduced by ILC, having m ore zeros than poles,

poles outside the unit circle, and sensitivity function gains other than unity. The

logarithmic average am plitude o f the sensitivity response functions is no longer

restricted to be zero or positive, and this generalized theorem gives one the ability to

determine how negative this average can be, i.e. how much overall attenuation is

accomplished in the ILC frequency response function.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
76

-0.5

-0.5 0.5

Fig. 4.1 The contour used in evaluating the Fig. 4.2 Convergence in an annulus — plot
integral o f (ln (l- z ) ) / z . o f R(z) when z = r e x p ( /0 ) for 0 < Q < k .
Solid line r - 1, dash line r = 0.9701, dash-
dot line r = 1.1499.

0.8

0.6

0.2

100 200 300


co
Fig. 4.3 Plot o f )| versus co up to Nyquist frequency.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
77

C hapter 5

T he W aterbed E ffect and R epetitive C on trol w ith Filtering

5.1 Introduction

Repetitive control is a relatively new field that uses control theory based

methods to leam to eliminate the influence o f a periodic disturbance on the output o f

a feedback control system, or to leam to elim inate determ inistic error in a feedback

controller in following a periodic command. The repetitive control law uses error

inform ation from the previous period to update the com m anded trajectory to the

feedback controller in the current period. The field normally considers one periodic

disturbance. This means that there can be a fundam ental frequency having this period,

and then there can be any number o f harm onics going up to Nyquist frequency. This

can form any arbitrary periodic disturbance w ith the given period.

One large class o f spacecraft uses a m om entum wheel for attitude stiffening.

A small imbalance in the wheel produces vibrations in the spacecraft structure, w ith a

fundamental period associated w ith the tim e for one rotation o f the wheel, and then

there can be harmonics. There are other classes o f spacecraft with one primary source

o f vibration, such as a cryo pump. W hen there is a fine pointing instrument on board,

it may be necessary to isolate the instrument from there vibrations. Passive methods

are simple, but may not give good enough perform ance. Typical feedback control can

do better, but such a system does not take advantage o f the knowledge that the jo b o f

the controller is to eliminate a periodic disturbance, i.e. one that can be predicted.

Hence, feedback control designs approach each new period o f the periodic

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
78

disturbance as if it had never seen it before. The result is to always be behind in

reacting to it, and not be able to totally cancel its effect on the output. Repetitive

control (RC) addresses this issue, and with sufficient design effort can entirely

eliminate the disturbance effect on the output.

The waterbed effect in feedback control says that if the control attenuates the

effect o f output disturbances in some frequency range, it m ust amplify disturbances in

some other frequency range [30]. This implies that feedback control works because it

attenuates in a range where there are significant disturbances, and amplifies in a range

where the disturbances are small. In chapter 4 and [32] the waterbed effect was

studied as it relates to iterative learning control (ILC). ILC differs from repetitive

control in that instead o f a periodic disturbance or command, the system is given the

same command repeatedly, but it is restarted from the same initial condition each

time. It is shown there how ILC is able to bypass the waterbed effect, and get better

performance than any feedback controller could do. ILC and RC are very similar,

w ith the m ajor difference being the restarting in the case o f ILC. It is the purpose o f

this chapter to study the relationship o f the waterbed effect to repetitive control.

5.2 Iterative Learning Control and Repetitive Control Background

In this section, we describe the same system in chapter 3 but represent in

slightly different form. The solution to equations (3.3) and (3.4) can be packaged in

matrix form as

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
79

y - p v M + w + ^ , „ X V( 0 ) ; ^ vo (5.1)

Linear learning laws take the form [18,19]

U.J+1 = ^ ( m 7 + ^ £ y) (5.2)

where L is the learning gain matrix, and F is a chosen filter which we normally use as

a zero-phase low-pass filter in this chapter, although we also consider non-zero phase

filters as well. Define the error e j = y —y ■ The above is given in terms of

matrices to describe the error each repetition. If each repetition is sufficiently long

that we can ignore the effects o f initial conditions, we can switch to a z-transfer

function representation, using G(z),L(z),F(z) for the plant, learning control law, and

filter in place o f the corresponding matrices P, L, F. Combining the above equations

produces the difference equation describing the error history as a function o f j. and

the sensitivity transfer function from command minus periodic output disturbance to

the resulting error (see [21 ])

Ej(z) - F(z)[I - G(z)L(z)]E/_l (z) = [1 - F(z)][F£/(z) - W(z)]


(5.3)
S, (z) = ---------- w ---------
l-£ (z ) [l- G (z )£ (z )]

Stability is a property o f the homogeneous difference equation, so we drop the forcing

function on the right when considering it. Then the coefficient matrix for

appears as a transfer function from the error in one repetition to the error in the next.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
80

If we ask that the associated frequency transfer function have amplitude less than

unity for all frequencies.

|F ( r ) [ l - G( z) L( z )]| < IVco,z = e,olT (5.4)

this suggests m onotonic decay o f the error with repetitions. Here. T is the sampling

time interval. The above developm ent is heuristic, based on steady state frequency

response thinking, but it can be shown that it is in fact a sufficient condition for

asymptotic stability o f the learning process [19]. In practice, without using a filter F.

it is hard to create a learning law L that is guaranteed to satisfy this inequality. So the

design process is to pick a n L that satisfies it up to some frequency, and then use a

zero phase filter F to cut o f f the learning above that frequency to obtain stability at

the expense o f not attem pting to cancel high frequency errors above the cutoff [19].

Now consider real tim e repetitive control. Converting equation (5.2) to the

transform domain, and recognizing that going back one repetition is now going back

p time steps produces the equivalent of equations (5.2) and (5.3)

z pU { z ) = F(z)\U(z) + L(z)E(z)]
[zp - F ( z ) + G(z)FCz)L(<z)]E(z) = [ z p - F{z)]\Ytl( z ) - W { z ) \ (5.5)
z p —F(z)
S r (Z) = z p - F{ z ) [ \ - G{ z ) L { z ) ]

and the homogeneous equation whose solution determines stability o f the process can

be written as z pE{z) = [/r (z)(l - G ( z ) L ( z ) ) ] E ( z ) . Interpreting the left hand side as

the error in the next period, we have the same heuristic argum ent for stability as

before, and again ask that equation (5.4) be satisfied. As before, it is possible to

demonstrate that this condition is in fact a sufficient condition for stability [19].

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
81

5.3 Batch Zero-Phase Low-Pass Filtering

Form f o r Im plem entation: N ow consider the form of the zero phase filter F used in

ILC. and also used in RC when the update is done as a periodic batch process. Pick a

causal low pass filter such as a Butterworth filter. In state space form

x(k +1) = Ax(k) + Bv(k)


(5.6)
v(k) = Cx(k) + Dv( k)

where we allow a D term which is needed for most such filters. If this is used to filter

a chosen length set o f data v(k), the result is to attenuate the am plitude o f frequency

components above the cutoff and leave the amplitudes below the cutoff close to their

original values, but considerable phase delay in introduced. Then the same filter is

applied again to the resulting sequence, but done in reverse order

x(k - 1 ) = Ax(k) + Bv(k)


_ (5.7)
v(k) = Cx(k) + Dv(k)

This doubles the attenuation o f frequency components above the cutoff, and cancels

the phase lag o f the previous filtering by a corresponding phase lead. A fter filtering

both forward and backward through a data set, the v{k) is the zero-phase low-pass

filtered value o f v ( k ) . One must pick initial conditions at time 0 for the forward

computation, and initial condition at the final time for the reverse computation. There

are various ways tc do this as discussed in [29], together with ways to extend the

trajectories to minimize the transient portions o f the filtered result.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
82

M atrix Form: Use the underbar as before to represent the whole history o f the

associated variable. Suppose that we are using the zero phase filter on a set o f data

points starting with time step ka and going through time step kj-. so we make the

underbar vector start with this first time step and progress to this final tim e step. The

one can write the solution to (5.6) as

v = V v + Aiex(k0) (5.8)

Here the P ,Aic are analogous to the Pv,Am in equation (5.1) w ith the A.B.C o f the

filter substituted and also with the influence o f D taken into consideration. This

means that the diagonal entries in P become D, and the other entries are pushed down

one. And the powers o f A in the Aic become one less than before. Let Ir be a matrix

like the identity matrix, but the opposite diagonal in the matrix contains the nonzero

entries. Then l rv_ puts the result o f the forward filtering from (5.8) into time reversed

order. Sending it through the same filter again, and then returning to the original order

produces

V= Ir[PIrY- + A,ck(kf )\
= + V rPIr)At x(.kJ + I,A,Akf) (x9)

One can show that / rP / r = P 7 for matrices o f the form o f P, so that the overall result

o f the zero phase filtering is

y = P r Py + P rAlcx(kt>) + IrAicx{kf ) (5.10)

If we want to pick the initial condition for the reverse time filtering as the final state

o f the forward filtering, it can be done by writing

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
83

x { k f ) = c [ A c j x ( k 0) + Prv]
C = [0 - 0 /]

where Alc t ,Pr indicate the matrices Ak ,P with the C replaced by the identity matrix.

When the trajectory is extended in various ways, one can develop the corresponding

formulas analogously.

Convolution Sum Representation: Denote the unit pulse response o f the forward

time filter, D,CB,CAB,CA2B,..., as /z(0),/z(l),/z(2),/z(3),.... Then the convolution sum

form o f equation (5.10) is

*/—*» kr~K,
+ k) = f U - k )vU + k„) + X W ~ k )CA'x (ko) + CAk,~K~kx(kf ) (5.12)
J=k i—k

k r - k 0- k kf -ka- j

f ( j \ k ) = X h( 0 K l + k ~J) for k > / ; f { j , k) = ^ h{i + j - k)h{i) for k < j


1=0 ,= 0

z-Transfer Function Representation: In making 2-transfer functions one sets the

initial conditions to zero, which drops the second and third terms on the right of

(5.10) or (5.12). The z-transform o f the unit pulse response is given by

H(z) = [C(zl - AY' B + D ] = h(0) + h(l)z~l + h(2)z~2 + /z(3)z'3 + ••• (5.13)

Examine the product

H ( z ~ l) H ( z ) = ^ T h 2(i)z° + ]£ h(i)h(i + l)z ‘ h(i + \)h(i)z '■ +


<=o ^ ,=o ,=o (5.14)
]T h(i)h(i + 2 )z 2 + 2 h(i + 2 ) /z ( /) z '2 + •- -
(=0 ;=0

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
84

Look at the P r P in (5.10) and write the product for a matrix large enough that the

h(J) get to negligible values. Equivalently, examine the f ( j . f c ) in equation (5.12)

and look at a tim e step k sufficiently far from the endpoints that the upper limits can

be considered independent o f k. j and can be replaced by infinity because the

elements h have becom e negligible. This makes / ( j . k ) become dependent only on

the difference j —k . The product o f P 7P with a unit input at a time step k

sufficiently far from the beginning or end o f the m atrix that end effects are not

important, produces an output that is the corresponding column o f P r P . This

corresponds to the infinite limits on f { j , k ) . Then exam ine the product H{z~x) H{ z)

and identify by the power o f z the output at different times, and one sees that they

match. Hence

V{z) = F'Xz)V(z) ; Fl ( z ) = H(z~x) H{ z ) (5.15)

The z-transfer function o f the zero-phase filter is Fr (z) when implemented in this

manner. Note that if there is a pole inside the unit circle in H(z), then there will be a

corresponding pole outside the unit circle in H(z~x). One might think that this

suggests instability, but in the finite time zero-phase filtering problem, the filtering is

noncausal, and poles outside the unit circle do not im ply a stability problem. Hence,

when studying the waterbed phenomenon for ILC with zero phase filtering we can

need the generalization to poles outside the unit circle given by the GBI Theorem.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
85

5.4 Real Tim e Zero-Phase Low Pass Filtering

Zero phase filtering is not a causal operation under normal circumstances, but

with a som ew hat heavy com putation burden each tim e step, we can create a real time

filter for repetitive control [20]. Repetitive control laws such as the linear phase lead

Law only require one zero phase filtered value each tim e step. A t time step n it wants a

zero phase filtered value o f u(n —p) + <pe(n —N + y ) , where p is the period o f the

periodic disturbance/com m and, 0 is a repetitive control gain, and the integer y is a

linear phase lead chosen to keep the frequency response plot o f (5.5). as used for

equation (5.4), satisfied up to as high a frequency as possible before using F to cut off

the learning. Let N = p —y be the number o f tim e steps back from the current time

step to the step o f the estim ated quantity. Then at tim e step n the forward computation

o f v(«) is as in equation (5.6). The needed backward zero-phase filtered value at time

step n —N can be written as a superposition o f previous v(k) according to

v(n —N ) = Cx(n —N ) + Dv(n —N )


= C A Nx(ri) + h(N)v (n —1) + h ( N —l)v(« —2) + (5.15)
— i- /z(Dv(/7 —jV + 1) + h(0)v(n —N )

Provided one can perform this sum within one tim e step, one can implement a real

time zero phase filter for repetitive control. This form ula ignores the initial condition

at time zero, since it will get arbitrarily far back in time as time progresses. But the

other initial condition used on the backward filter starting back at time step «, x ( n ) , is

included in the equation in case the number o f time steps N o f decay is not

sufficiently large to make the term negligible.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
86

To create a transfer function representation for (5.15). we ignore the first term

and denoting n —N by k . create the z-transform o f the corresponding sequence.

Then

K(z) = [h( 0) + K l ) z + h{2)z2 + - + h( iV) z v] [C( z / - A)~l S + D]V(z)


(0 . 1 6 )
= Hy (z-l)H(z)nz)

This differs from the previous result (5.15) because the H(z~x) expansion has been

truncated to N terms. Note that the / / v(z_l) only introduces zeros in the transfer

function, and hence the poles o f this zero phase filter are simply the poles o f the

forward low pass filter. This time there will be no poles outside the unit circle, as is

required for stability in a real time process.

One way to pick the initial condition for the backward filter is to make its

initial state equal the final state o f the forward filter, x(ri) = x ( n ) . Doing this and

recognizing the time argument k being used, produces the different z-transfer

function

V{z) = \ z s CAs { z l - A )'1B + H s (z -‘)H(z)]V(z) (5.17)

Still different version would result if we chose to extend by repeating this state or by

doing an anti-symmetric extension around the final value [29].

5.5 Some W aterbed Conclusions

In learning and repetitive control, we can apply the Bode or generalized Bode

integral theorem to two different transfer functions, (i) the appropriate sensitivity

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
87

transfer function or (ii) the transfer function from one repetition to the next

F (z)[l —G( z ) L( z ) \ .

Learning Control: Reference [32] gives exam ples o f this case. First consider ILC

laws that converge to zero error, and do not need a cut o ff filter making F( z) = 1. In

the difference equation in (5.3), the right hand side is then zero, m aking the

sensitivity transfer function zero. The desired trajectory Yd(z) and the disturbance

W(z) repeat every repetition. Every repetition contains p time steps limiting the

number o f frequencies that can be observed to only those that are periodic with period

p, and there is not way to have any other observable frequency in the system. With

the forcing function zero in difference equation (5.3), a sufficient condition for the

error to go to zero is (5.4). For (5.4) to be satisfied, the GBI Theorem m ust produce a

negative value. Suppose that the learning law was the simple integral control based

ILC having L(z) = (frz. Examining the transfer function for (ii) above we see that the

coefficient o f the highest power includes the learning gain <f>, requiring the use o f the

generalized Bode integral theorem, and allowing the integral to be negative because

o f the first term on the right in (4.7).

N ow consider using a zero phase low pass filter FL(z) and a learning control

law containing a linear phase lead, L(z) = <pzy w ith any integer y > 1 corresponding

to a phase lead that can help keep equation (5.4) satisfied up to a higher frequency

before the low pass filter has to cut o ff the learning. N ow the right hand side o f the

difference equation (5.3) is non zero and there is a sensitivity transfer function (5.3)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
88

to consider. The standard Bode integral theorem does not apply for various reasons.

Increasing y above one makes more poles than zeros in the sensitivity transfer

function. It also allows one to have a non unity value for the gain K r in equations

(4.7.4. 8 ). And in addition, since we are using a noncausal zero phase filter, there will

be poles outside the unit circle. Hence, ILC can bypass the waterbed effect in the

steady state error levels reached.

Again, for stability we would like to satisfy (5.4), and this means that we m ust

be able to bypass the standard waterbed effect again for the transfer function o f (ii)

above. This is accomplished in several ways. The poles outside the unit circle in the

zero-phase low-pass filter help decrease the integral according to (4.7). Also,

increasing y above one makes the coefficient o f the highest power o f r in the

num erator contain the learning gain <f> as a multiplicative factor making the first term

in (4.7) help.

Figure 5.1 plots the sensitivity function for an example system. We use the

same exam ple system for essentially all numerical results presented. It is a m odel o f

the closed loop transfer function for each link o f a Robotics Research Corporation

robot. In continuous form the transfer function is o f third order as given in (3.16)

where d = 8.8 rad/sec, co0 = 3 7 rad/sec, and £ = 0.5. A sample rate o f 100 Hz is

considered, and the low pass filter is a 5th order Butterworth with a cutoff at 100

rad/sec equivalent to 15.9 Hz. Using the filter in zero phase form gives a 10th order

attenuation. The learning gain tf>= \ and the phase lead parameter y = 5. It is clear

from Fig. 5.1 that the learning control law has very effectively bypassed the waterbed

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
89

result. The plot goes to Nyquist frequency, and although it is very slightly above unity

for frequencies above the cutoff, there was no significant penalty paid for the good

performance below the cutoff. N ote that the sensitivity function and the GBI Theorem

do not have any information about the finite num ber o f time steps p in the ILC

problem. They talk about steady state frequency response, and Fig. 5.1 gives results

for all frequencies. Which points on the plot refer to the problem at hand are

determ ined by the frequencies that can be seen in p time steps o f data. For large p

there will be many points, and for small p there are a small number o f frequencies. O f

course the plot is for steady state response, and therefore only applies for sections o f

the p step trajectory after initial condition influences have become negligible.

Real Time Repetitive Control: N ow consider the sensitivity function SR o f equation

(5.5) for real time repetitive control. This sensitivity function differs from the

previous S i in that the Ts are replaced by z p . This fact prevents a phase lead in the

learning law from influencing the highest power o f z in the denominator. It would

require that y be larger than p , and this means we would need a future error. Also,

the zero-phase filter FL(z) m ust be replaced by the causal filter FR(z). This precludes

having poles outside the unit circle. The conclusion is that when using real time

repetitive control one is necessarily subject to the standard waterbed effect of (4 .5)

concerning the sensitivity transfer function of equation (4.5).

N ow consider the transfer function of (ii) above. It does not really apply as a

transfer function from one period to the next, except when the learning is sufficiently

slow that one can make a quasistatic assumption. Nevertheless, we are still interested

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
90

in satisfying equation (5.4) for stability, and this requires that the transfer function

involved bypasses the standard waterbed result. Both causal and zero phase filters can

be used in satisfying this stability condition. Increasing / above one, helps by

producing the gain term on the right in the GBI Theorem (4.7). In the case o f a real

time zero phase filter, there are no poles outside the unit circle, but the H x (z~l)

introduces many zeros, creating more zeros than poles.

Long Term Batch Update Repetitive Control: So far the discussions o f repetitive

control in this chapter have only considered the implementation with real time control

updates. An alternative is a long-term batch update o f the command to the feedback

controller. Suppose we use the learning control laws as discussed above in a long

term batch update mode. The repetitive control problem can have a constant

command and a periodic disturbance o f period p time steps, or it can have only a

periodic command, or a periodic command and disturbance both o f the same period.

First, apply the desired output as the command to the feedback controller, and wait

until the response is considered to be in steady state. Take a substantial data set o f this

steady state response, and compute the zero-phase filtered value and the learning

update. Then apply the result, wait for steady state, make the next update in the

command, and continue the process. Since all responses used are considered to have

reached steady state, the frequency response information in the sensitivity transfer

function (5.3) with z = exp(icoT) , is correspondingly more accurate than in the

learning control case. The possible repetitive control laws that can be used when

implemented in this batch update mode, can be the sam e as the range o f learning

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
91

control laws. A nd with sufficiently long waiting between each update, one should be

able to get arbitrarily close to the steady state response indicated by the learning

control sensitivity transfer function, e.g. Fig. 5.1. We will dem onstrate that this is the

case later (Figs. 5.17, 5.18).

However, the above thinking only applies to frequencies that are periodic with

period p tim e steps. Unlike the learning control problem w here one only has p time

steps o f data and hence there are no frequencies in the problem that are not

expressible w ith period p, in the repetitive control problem w e can have deterministic

disturbances that are periodic with som e unrelated frequency. The long term updates

must be tim ed to coincide with the phase o f the disturbance/comm and o f period p

time steps. Any unrelated determ inistic frequency will produce a learned response

whose phase could be at worst 180 degrees o ff in the next application o f the learned

corrective signal. And this can produce errors that can double the original error size.

This will be demonstrated later (Figs. 5.13-5.16), and the underlying mechanism is

similar to the discussion of Figs. 5.5 and 5.6 below. Although one can get

amplification o f errors in this batch update regime, the Bode Integral Theorem type o f

thinking does not apply to the problem. As formulated below, one needs a time

history o f the output over one or m ore periods for batch update vectors, and the

frequency response is seen as the frequency observed going down the entries in the

column vector. In addition, there are start and end effects in the colum n vectors, so

that the output is not a single frequency coming out for a single frequency going in. A

m athem atical w ay o f predicting the response is developed later.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
92

5.6 The Need for Zero-Phase Low-Pass Filtering in RC

From the point o f view o f creating guaranteed stability using equation (5.4).

there is not distinction between using a causal low pass filter o f a zero phase low pass

filter F(z). However, the performance o f the system according to the sensitivity

transfer function (5.5) can be very different. Figure 5.2 plots the steady state

am plitude frequency response o f this sensitivity function when a 10th order causal

Butterworth filter is used with the third order feedback control system and linear

phase lead learning law discussed above. Figure 5.3 is the corresponding plot using a

real time zero phase 5th order Butterworth filter, which is effectively 10th order

because o f the combination o f forward and backward filtering. These plots go only up

to coT = 1.5 , but the remainder o f the plots up to Nyquist at n hold no surprises.

These plots look very similar, and both show the possibility o f amplifying by a factor

o f more than 5 at some frequencies different from those periodic w ith period p.

However, looking at where the curve approaches the horizontal axis near the point

coT = 0.5 one sees that the minimum points are different. In Fig. 5.2 the frequencies

that the system is trying to eliminate having period p are indicated by circles, and it is

clear that the causal filter does a poor jo b o f attenuating at these desired frequencies.

Figure 5.4 plots ju st these frequencies for both cases, and it is clear that even for the

fundamental o f period p the causal filter does not attenuate as much as 50% o f the

error, and that for the first harmonic it has actually made the error worse by a factor

2.3. This is easily understood by looking at the frequency transfer function for the

numerator in sensitivity transfer function (5.5) z p —F{z) = e ,blVp - M{co)e,0{OJ). For

small co the M{co) is close to unity, and this function goes through a m inim um near

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
93

zero when the phase angles match, i.e. when a> = \ 2 x /(Tp)]l + 0{co) /(Tp) where i is

any integer. For the zero phase filter the 9(co) is zero, and this means the minima are

at the desired frequencies o f period p , but when 8{a>) is not zero for the causal zero

phase filter, the minima occur at the wrong frequencies. Since the minima are narrow,

being somewhat o ff can produce amplification o f frequencies that were supposed to

be attenuated.

The above comments apply to the situation when the low pass filter is used as

in (5.4) to produce stabilization. In [13] a causal filter was used on the error only, and

stabilization was accomplished instead by quantization. The linear phase lead was

used to compensate for both the system phase and the filter phase, and good

performance was obtained. A batch zero phase low pass filter combined with this can

periodically eliminate any transient accumulation effect in the final error levels.

5.7 Understanding The M echanism for Amplification

A simple example can give understanding o f both how amplification is

accomplished, and at the same tim e show some limitations o f the Bode integral

thinking. Again examine the numerator o f the sensitivity transfer function (5.5), but

for simplicity supposes that we are able to satisfy (5.4) without needing low pass

filtering, so the numerator is z p - 1 . The magnitude frequency response plot of this

term is given by \[e'“>Tp - l j = [2 - 2 cos( 6>7/ 7)]‘' 2. For the frequencies periodic with

period p, this give the desired zero value. For frequencies half way between looking

back p time steps is looking back an odd number o f h alf periods o f the oscillation, so

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
94

that the terms adds the signals rather than subtracting th em . Thus the plot shown in

Fig. 5.5 (shown up to tt! 5 but the behavior remains th e same to Nyquist), and the

plot goes from the m inim a that cancel the frequencies o f interest to m axim a that

amplify by a factor o f 2 the frequencies one h alf w a y between the harm onics o f

period p. N ow exam ine what the average is over all frequencies up to Nyquist:

(1 /tt) f | txp(i9p) \dQ = 1.2732. This means that the te rm amplifies the disturbances
JO

when averaged over all frequencies. However, the Generalized Bode Integral

Theorem applies to this transfer function, and it gives th e average o f the logarithm as

zero. From this one m ight presume that there was no overall amplification on the

average, which we know is incorrect. The source o f the difficulty is that the integrand

in the GBI Theorem contains a logarithm o f a function that repeatedly goes to zero,

making the integrand go momentarily to negative infinity.

Figure 5.6 gives the corresponding plot when a re a l time zero phase low pass

filter is used (employing a 5th order Butterworth w ith a c u to ff at 100 rad/sec). Now in

place o f moving back and forth between zero and two, th e amplitude o f this decreases

as the value o f M(co) decays from unity. The average o f z p —F{z) over all

frequencies up to N yquist is 1.080 for the zero-phase filter case, and 1.087 for the

causal low pass filter case (10th order). The zero-phase filter gives the better value,

but the difference is not large. The big difference is th a t the causal case put the

minima in the wrong place. The corresponding Bode integral values are 1.006 x 10“ 13

for the zero-phase filter case, and 2.509 x 10-14 for ther causal filter. This time the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
95

better value is for the causal filter, but this difference might be at the numerical noise

level. Both values are essentially zero.

N ow consider the question o f how the amplification show n by the sensitivity

function becomes larger than 2. for the peaks in Figs. 5.2 and 5.3. A t low frequencies

the factor [1 - G(z)L(z)] w ill be small, and can be near zero if the DC gain o f G is

unity and the learning gain is unity. Then one would observe am plification by a factor

o f approximately 2, as observed in Figs. 5.2 and 5.3 for the first peak. At high

frequencies the F in the num erator and the denominator goes to zero, making the

sensitivity transfer function approach unity at all frequencies in this range. For

intermediate frequencies before F has become small, the square bracket term in the

denom inator will introduce a phase change. For G having a phase lag as the

frequency goes up, the [1 —G(z)Z(z)] will exhibit a corresponding phase lead. The

denom inator will go through a minimum around where the phase o f z p matches that

o f the bracket term (for F zero phase), and this shifts the peaks o f Fig. 5.3 toward the

early side o f the intervals between harmonics o f period p. A nd o f course, the extra

am plification above a factor o f 2 at these points occurs because the denominator can

become small in magnitude.

5.8 Comments on Stability Conditions and Stability Boundary

Sufficient Condition f o r Stability in RC: Equation (5.4) has been described as a

sufficient condition for stability. Here we summarize the discussion in [20]

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
96

establishing this for L(z) = (fc.Y. The characteristic polynomial for (5.5) is the

num erator o f the square bracket term on the left o f this equation. Suppose the

denom inator has only roots inside the unit circle (G and F are asym ptotically stable).

We could establish stability according Nyquist methods by m aking a contour going

counterclockwise around the unit circle, out to infinity along a branch cut, around at

infinite radius clockwise, and back along the branch cut. Then if the phase angle

comes back to its original value at the end, there are no roots outside the unit circle.

Instead o f using the term in square brackets on the left o f (5.5), we could use

z~pF(z)\\ —G(z)L(z)] and see if the phase angle measured from +1 comes back to

its original value to determ ine stability. This new expression is strictly proper

provided y —1 < p when G has one more zero than pole which is generically the case,

and y < p when the number o f zeros equals the number o f poles in G. Note that these

inequalities correspond to not asking for a future error in the learning control law.

U nder these conditions, the phase angle condition cited above defines the true

stability boundary, as a necessary and sufficient condition for asymptotic stability.

Note that for z on the unit circle, the magnitude [z^j = 1, and then satisfying (5.4) will

automatically satisfy the necessary and sufficient condition developed here. The

corresponding results showing sufficiency o f the frequency condition (5.4) for

learning control, and examination o f the distinction between this sufficient condition

for stability and the true stability boundary are discussed in detail in [19,31],

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
97

Testing f o r the Stability Boundary Without a Filter: The normal Nyquist plot for

(5.5) is made looking at the characteristic equation written in the form

F W C K xW r)
z'-fW

When there is no low pass filter, F(z) = 1 , and the denom inator o f (5.18) contains p

roots on the unit circle. The contour m ust be adjusted to go around each o f these roots

when making the Nyquist stability analysis. This difficulty can be avoided by using

the method o f [28] as done in [31, 15] which instead plots 0 ( z ) = (1 —z p) / ( z yG ( z ) ) .

The equation (2.4), Z = (- W / 180°) + Pq + n ' I 2 , determines stability. Pick any point

(p on the positive real axis and track the change in phase angle from this point to the

points Q{e'°) as 9 goes from zero to 180 degrees, and this is W. The Pq is the

number o f poles of O outside the unit circle, i.e. for this application, the number of

zeros o f G outside. The n' is the num ber o f poles of O on the unit circle, which is p in

this application. Then Z is the num ber o f zeros o f 1 + 0 outside the unit circle, i.e. the

number o f unstable roots o f the characteristic equation.

Introducing a Filter F: When a Butterworth low pass filter F is introduced the term

z p - F(z) no longer has roots on the unit circle except at z = 1, since |/r(e "9)| < 1 for

all 0 < 9 < 180°. But this makes the usual Nyquist plot singular at zero frequency

which is still inconvenient in the plotting process. An alternative is to plot O again.

However, the Butterworth filter has a factor (z + 1)'" for its numerator, where the m is

the order o f the filter (or two times the order o f the filter in the batch zero phase case

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
98

with two applications o f the filter). There for by using O, the singularity has been

switched from zero to N yquist frequency, and again it becom es hard to plot and count

encirclem ents.

Assum e that the unit circle condition (5.4) is satisfied by for repetitive control

gains in the range 0 <<f> < $ max . T hen consider plotting O. One has trouble finding W,

but one knows Pq and n' , and this implies that we know W for the above range of <fi.

The range o f stabilizing (f> will continue past the value (pxmax, which is a sufficient

condition for stability, until it hits the actual stability boundary <f>2max. And this new

gain is obvious for the beginning o f the plot of Q, giving the interval on <p up to the

first loop o f the curve. Hence, this forms a simple m ethod to answ er the following

question. Pick a repetitive controller that satisfies the sufficient condition for stability

(5.4) for sufficiently small gain (f>, and then ask how far beyond the maximum

allowed gain satisfying (5.4) can one go before actually reaching the true stability

boundary? This approach allows one to adjust the gain to reach the stability boundary,

but it does not allow one to adjust the cutoff frequency upward violating (5.4). Figure

5.7 shows the 0 Ijnaxand f ° r the third order system (3.16) using a normal causal

10th order Butterworth filter w ith a 100 rad/sec cutoff and a sam ple rate o f 100 Hz.

The control law as before uses a phase lead parameter o f y = 5. This time the

parameter con is allowed to vary along the horizontal axis. The sufficient condition

(5.4) produces maximum values o f the gain that are very close to the true stability

boundary. This is consistent w ith results in [31].

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
99

Stability f o r Batch Update Repetitive Control: In order to make this stability

discussion address each type o f problem, we com m ent on the stability o f batch update

repetitive control as discussed in the next section. W hen updates are made every p

time steps, stability is determ ined by seeing whether all eigenvalues o f the matrix A

o f equation (5.20) below have magnitude less than one, and for the case o f updates

every 5p one uses the A o f equation (5.30) below. Using p = 100 and the third order

system o f (3.16), the m axim um eigenvalue in the former case is 0.9680 and in the

latter case it is 0.9725.

5.9 Frequency Response o f Periodic Update Repetitive Control

The developments o f the previous sections say that if the repetitive control

law does an update in real time, it will be subject to the limitations o f the waterbed

effect. On the other hand, iterative learning control with zero phase filtering is not

subject to this effect. By m aking batch updates in repetitive control, and by waiting

until reaching steady state before making an update, one expects that one can obtain

the same performance in repetitive control as in learning control, at least for tracking

errors o f period p. In this section we investigate what the final error levels should

look like for this mode o f operation, for the frequencies periodic with p steps, and

then we look at the behavior at other frequencies. We observe that amplification by a

factor o f 2 is again possible, but by doing batch updates, amplifications above this

value are no longer likely to occur in the steady state response. It takes some

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
100

mathematical development to create the equivalent for batch update o f the frequency

response o f the sensitivity transfer function, and this is the purpose the section.

A M atrix Formulation f o r Updates Every Period: First consider m aking a batch

update at the end o f every period. This is a limiting situation that m ight require some

fast com putation in the last time step for the zero phase filter. In the next section we

will relax the computation requirement. Using the underbar notation for the history o f

variables in a period as introduced for equations (5.1) and (5.2), we can write the

relevant equations as

£j = + A mxj(0)
y j = C X j + w; xM (0) = CXj (5.19)
u J+ 1 = F ( U j + L e j )

For sim plicity both the forward and the reverse time initial conditions for the filter F

are set to zero for this development. Matrices P r,Am are P r,Am with the output

matrix C v removed, and the matrix C is a block diagonal matrix with C v along the

diagonal. For this repetitive control problem the first entry in x } is the initial

condition, and according to the third equation o f (5.19) the initial condition for period

j + 1 is equal to the final state for period j . A minimal state vector to use for the

periodic updates must contain the p step history o f the inputs as contained in and

the initial condition for each period x y( 0 ) . From this state xj the first o f equations

(5.19) gives the complete state history for period j as an output equation.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
101

Developing the update equations for the two parts o f th is new state vector and

packaging them together produces

x=y+1 = A= x= / + = y - w—)'
B (_«/ (5.20)

r c
——a zo
xo C
--- Pr
x 1 ' 0 '

;A = :B =
-F L C A xn F (I —LCPX)_ - F L .

Suppose that y - w is constant so that the command a n d the disturbances are both

periodic w ith period p tim e steps. Then when steady s ta te response is reached,

x j+i = x j , and we can solve for the corresponding steady s ta te error as

!„ = { /-C k „ -K ) (5.21)

This is a matrix version o f the sensitivity transfer function. There is no complete

analog o f the frequency transfer function for such a batch: update process. W hat we

want to know is the frequency content o f the history o f tfie error that one observes

going down the column matrix e ss. We can substitute a sin u so id for either the desired

trajectory or the output disturbance, and obtain the response. This process works for

all sinusoids that produce constant values for y or w Tor all repetitions, i.e. for

frequencies that have p time steps as their period, a fundam ental with that period and

all harmonics up to Nyquist. Since the desired trajectory is supposed to have this

period, this equation is sufficient for studying the tracking ezrror to commands.

On the other hand, it is o f interest to know how t h e repetitive control system

behaves when confronted with a disturbance that does not h ap p en to be at one o f the

frequencies targeted for elimination by the repetitive controller. It is not obvious how

to do this for all frequencies, but we can generate the frequency response information

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
102

for a countable infinity o f frequencies by the following procedure. The procedure is

illustrated for tripling the num ber o f frequencies addressed. To plot two points

betw een each o f the ones obtained above, consider frequency com ponents in the

output disturbance that are periodic with period 3p. This can be represented w ithin the

current framework by using a w, for period j , a w, for period j + 1, and w3 for

period j + 2, and repeat for J = 1,4,7,10,... This produces the dynamic equations from

one o f these values o f j to the next

x j^ . = A x j + B y j - w 2 ;A = AJ ' , B = A ZB AB s] (5.22)

To develop the associated steady state error, first find the steady state value o f the

state vector for these j , by setting x j = x ^ , = x v[, and then propagate this state to the

next two periods

; lv.v. = { y d - Vf,) - C[ a „ Pr ]xvvi (5.23)

Then the steady state error history is periodic with 3p, and is given by e for

j = 1.3.7.10,... and by | lt, . | tv3 for j + 1 and j + 2 respectively.

Consider the third order system again, and consider a period p o f 100 time

steps. To study the final error level associated with the 5th harmonic at 5 Hz. set the

command (5.21) to equal the sampled version o f sin(lOzzr). The resulting steady state

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
103

error is plotted in Fig. 5.8. The corresponding plot for the 14 H z harmonic is given in

Fig. 5.9. A linear phase lead param eter y o f five steps was used. In matrix form for

the batch update, this means that there is no learning in the last four steps o f the data

set used, and this can be the cause o f some o f the end effects observed. The more

important consideration is the zero phase filter initial conditions at the start and the

end o f this time period. It is clear that one needs to pad the ends to allow time for

decay o f these initial condition effects, so that the choice o f their values has

negligible influence.

A M atrix Formulation f o r a N atural Update Rate: A natural way in which to

intersperse the batch update computation among the real tim e computations was

presented in [13]. This com putes one time step o f filtering update o f the batch zero

phase filter during each time step o f the real time process. To allow substantial time

for the effect o f filter initial conditions to decay, the forward filtering is started one

period in advance o f the period that is to be zero-phase filtered. Similarly, in order to

allow substantial time for the filter initial conditions to decay for the reverse path, an

extra period is added on the end as well. Therefore, the batch update has the

following features: the filtering is started at the beginning o f some period f. , and it

continues filtering the appropriate function as the data arrives through three periods,

including periods, I , t + 1 and t + 2. Starting with the 4th period, the results o f the

forward filtering are sent through the filter in reverse order, one time step o f the

forward filtered result is processed for each time step in real time. Once two more

periods have elapsed, the desired zero phase filter result is available for the period H

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
104

+ 1, the result containing transients for £ + 2 is discarded, and there is no reason to

continue the filtering through period £. This creates a batch zero-phase filter update

every 5 periods.

The subscript j is now reserved to count the batch updates, so that it increases

one unit every 5 periods. The Xj (0) is the system state at the start o f these 5 periods.

The vectors giving the history o f the states for each o f these 5 periods are

x jl,xj2r...xJ5, and the repetitive control action that is applied is the

same p step history for all 5 periods. Following the method o f equation (5.29) the

vectors o f the state histories are generated according to

Xji = P ,w 7 + Axitx / (0)


Xjj+i = p , “ y + A ^ C x j , :i = 1,2,3,4 (5.24)
xJ+i(0) = C x JS

As before the minimal state vector needed to know what happens every repetition is

that given in equation (5.20) containing an update of the initial system state at the

start o f the 5 periods, xy(0), and the repetitive control action, u] . Cascading the

equations in (5.24) produces the update for the first o f these two

Xj . l (0) = A uxJ(0) + An uj
(5.25)
du = ( Q A,ay ^ d n = U + C A „ + (£ A a,)* + (C T t„)3 + ( C 4 „ ) 4]

Now we need to develop the update equation for uJ+l. The zero phase filtering is

done on three periods o f the error, so we define e} to contain the error history for the

first three periods o f the five periods for update j, and it can be written as

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
105

Li *7.
- c (5.26)
*7 = L, -W j 2 *72

L, 7^ _

Here we have allow ed the disturbance for each o f the periods to be different so that

the final equation can be used to analyze disturbance o f any period. O f course, when

the disturbance contains only period p , then all o f the w are the same and can be

replaced by vv. During the first three periods, a causal zero phase filter is being

applied forward in tim e. Denote this filter by its matrix o f M arkov parameters P~ (we

let the initial condition be zero, since we are allowing a full period to the initial

condition effects to decay). Then the result o f the forward filtering o f the three

periods can be written as

“j
P3{ “ j (5.27)
y.j

Note that the learning law L is being applied for all three periods to the error history.

M ultiplying this by P.; produces the backwards filtering o f all three periods. The

product P.r P3 is the zero phase filter F. Then it is the center period that is o f interest,

and we can pick this period out by multiplying by the m atrix l m = [0 / 0]

producing

Hm = L F i Hj +Le j } (5.28)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
106

Com bining equations (5.24), (5.26), and (5.28) and packaging the result w ith equation

(5.25) produces the state equation for the repetitive control batch update

0
x
— j-r I
= ,4x + Ba y■Ld —-wJ - \B
---------- j — ’ =0 = (5.29)
v —— w j>~_ J « FL.
JLd

CAm ~i 1
CAxo(CAxo) ^ 2 2 = I » F 1 - K .F L C I + AxnC
CAxo(CAxu)2 1 J + Axoc + (A xoc ) \

and the A u ,Ar are given in equation (5.25), where these are the four partitions of

matrix A . W hen studying disturbances o f period p , this equation reduces to

xjrl = A x j + B ( y j -yv)-,B = Ba[l I if (5.30)

From these equations one can find the steady state error for disturbances of

period p , and also for disturbances o f period 5p, immediately. Consider the second

case containing the disturbance sequence w,,vv2,w3,vv4,u ’5 that occurs for the 5

periods o f each update. In steady state x =x = and

x = { I - A Y xBa (5.31)

This can be used in equations (5.24) to produce the steady state values for the five

periods x vvl ,x u2,x „3 ^,,5 >and then the steady state errors for the five periods are

= y d - C x „ , - w, ; / = 1,23,4,5 (5.32)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
107

Setting all w, to w gives the equations for determining the final error level in

response to disturbances o f period p.

Figure 5.10 shows the steady state error for the fundam ental frequency o f I

Hz. The error is quite small, but it does contain not only the fundamental but also

some harmonics. Figure 5.11 does a DFT on the 500 points and shows the harmonic

content. Figure 5.12 gives the result for a 2 Hz input, and the presence o f harmonics

is no longer obvious.

It is o f interest to understand the relationship between the final error level and

the learning gain. Reference [27] looks at the steady state covariance equation and

suggests that in the presence o f w hite noise the final error level will get better as the

learning or repetitive control gain gets smaller. The action taken by the learning law

as a result o f the noise will be smaller. Simulation here gave the opposite result for

the deterministic frequency components o f period p. For the fundamental in Fig. 10

the amplitude is about 6 x 10-13 using a gain o f <p = 1, and when the gain is decreased

by a factor o f 10 the amplitude increases by a factor o f 10. To explain this, examine

the learning control sensitivity transfer function, St = (1 —F) /(I —F + FGL) , for low

frequencies. In this range F is near unity and this means that the denominator is

dominated by the term FGL, whose value is approximately equal to 0, and hence

decreasing this gain increases the final error level in this range. This suggests that

final error levels when using a cutoff filter has a tradeoff in the choice o f gain for

final error level with respect to the performance for terms periodic with p and

performance in response to white noise.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
108

N ow consider what happens to the final error level w hen there are

determ inistic disturbances at frequencies other than those o f period p. One can o f

course apply the method outlined in equation (5.23) to the 5p approach in this section

in order to compute responses for as many intermediate frequencies as desired.

However, because o f the 5p computation in equations (5.29) and (5.31), we already

have the ability to examine four evenly spaced frequencies within the one Hertz

intervals between harmonics o f period p. Making use of these results. Fig. 5.13 shows

the steady state error for a sine wave input at 1/5 Hz, and also for a cosine input o f the

same frequency. Note that the error during the second one second period is nicely

cancelled, but the algorithms assumes that this same error will appear for all 5 o f the

one second periods, and therefore does inappropriate things for the other 4 seconds.

W hen the reaction is exactly wrong one expect to reach a m axim um error that

amplifies the original error by a factor o f 2. The mechanism involved is a generalized

version o f the mechanism discussed for Figs. 5.5 and 5.6. Figure 5.14 gives the

corresponding plots for sine wave inputs o f frequency 1/5 and 4/5 Hz. At 8 Hz the

coT = 0.502 which corresponds to starting up the knee o f the curve in Fig. 5.1. The

performance for this frequency that is periodic with period p is still relatively good

with an amplitude o f 5 x 10- 3 . For a frequency o f 8.2 Hz the steady state error level

to a sine wave input is shown in Fig. 5.15. The frequency 8.8 Hz produces a very

sim ilar plot, while 8.4 and 8.6 Hz produce plots o f the same nature but this time the

amplitudes o f oscillation in the first and third second are larger than the amplitudes in

the fourth and fifth seconds. Figure 5.16 gives the corresponding result above the

cutoff where the learning process is much attenuated, at the frequency o f 20.2 Hz

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
109

which is associated with coT = 1.268 . The response looks like the input except that it

is superim posed on a slow small am plitude oscillation. Only the tim e from 1 to 2

seconds is shown, but the remaining 4 seconds appear like a simple continuation o f

the same patter. For 20.6 Hz the behavior is sim ilar except that instead o f one small

amplitude oscillation in the one second shown, there are 3 such oscillations.

N ow exam ine the comparison between the performance o f ILC as seen in the

sensitivity transfer function o f Fig. 5.1, and the performance obtainable in long term

batch updates in repetitive control. To w hatever extent the long term batch updates

correspond to steady state frequency response updates associated w ith the transfer

function from update to update used in (5.4), the behavior should be the same as that

in Fig. 5.1. Actually, it should be possible to have Fig. 5.1 be m ore precisely

applicable to the batch update repetitive control than it is to learning control. In the

repetitive control problem time continues indefinitely allowing the system to

approach steady state frequency response behavior, while the learning control

problem is always limited to p time steps. O f course this perform ance is only

appropriate for frequencies o f period p time steps, so we compare these frequencies.

Equation (5.21) could be described as the sensitivity matrix for Ip updates analogous

to the sensitivity transfer function SL , and using equations (5.31) and (5.32) gives the

5p analog. Using this second analog, taking the discrete Fourier transform o f the 500

points in the error, and computing the difference with Fig. 5.1 produces the 5p curve

in Fig. 5.17. Agreement is quite good, with the peak o f the difference being

4.5 x 10 We can improve this agreement by allowing more time to reach steady

state response between updates, by following the same algorithm but using 2p in

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
110

place o f p. Figure 5.17 also displays the result for this case, but one cannot see the

actual size o f the error. Figure 5.18 plots the result w ith a different scale, and we see

that the amplitude the difference in performance betw een the learning control result in

Fig. 5.1 and this longer term batch update is roughly 6 x 10~15.

5.10 Conclusions

Repetitive control aims for zero error for the period associated with an

im portant disturbance source, at least up to some cutoff frequency for the learning.

This means that the frequency transfer function from disturbance to error m ust go to

near zero at the fundamental frequency and at integer multiples up to the cutoff. If the

waterbed effect applies, this attenuation periodically throughout a large frequency

range, perhaps up to Nyquist, must be paid for by amplification o f frequencies that

occur between these harmonics, or above the cutoff. It is shown here that real time

repetitive control is subject to this waterbed effect, and that the response can be quite

sensitive, going to zero for a chosen frequency and being amplified by a factor o f 5 at

a very nearby frequency. It is shown that batch update repetitive control can bypass

the waterbed effect for frequencies o f period p , approaching the same sensitivity

transfer function behavior observed for iterative learning control. For other

frequencies between those o f period p, there can be amplification, and the mechanism

for this amplification suggests that it is essentially limited to amplifying by a factor o f

2. This is in contrast to the real time repetitive control situation where the

denominator o f the sensitivity transfer function can go through a minimum producing

peaks that amplify by a factor o f over 5. Thus, long term batch updates have better

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Ill

steady state performance than real time repetitive control. This improved performance

is in exchange for slow er reaction times to changing conditions.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
112

55 0.8

0.6

0.4
CO

0.2

0.5 1 1.5 2.5


coT
a)
Fig. 5.1 Amplitude o f frequency response o f SL(z).

0.5
coT
Fig. 5.2 Amplitude o f frequency response o f SR(z) using a causal low pass filter.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
113

CO

coT
Fig. 5.3 Amplitude o f frequency response o f SR(z) using a zero phase low pass Filter.

2.5

>

o
<o
*o
3
■<— *

*5
on
CO
0.5

2.5
coT
Fig. 5.4 Comparison o f Figs. 2 and 3 at frequencies periodic with period p steps.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
114

0.5

0.2 0.3 0.4 0.5 0.6


coT
Fig. 5.5 Frequency magnitude plot o f d3- 1 to tc/5.

0.5

0.5 2.5
coT
Fig. 5.6 Magnitude plot o f \^-F(z)\ w ith real-time zero-phase filter

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
115

2 . 5,
N y q u ist sta b ility boundary for p = 1 0 0

A pproxim ated m on oton ic decay boundary

0.5

0 i
0.32 0.34 0.36 0.38 0.4 0.42 0.44
C0o T
Fig. 5.7 Comparison o f true stability boundary and monotonic decay condition.

0.5

-0.5

0.2 0.4
0.6 0.8
Time
Fig. 5.8 5 Hz command y j and steady state error with batch zero phase update every p
steps.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
116

0.5

-0.5

0.2 0.4 0 .6 0.8


Tim e
Fig. 5.9 Same as Fig. 5.8 for 14 Hz command.

o
fc
W
GO
GO

Time
Fig. 5.10 Steady state error for 1 Hz command, zero phase batch update every 5p.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
0 *
*1 " T S " "T " 1 ' ill>
Frequency
Fig. 5.11 Frequency content o f Fig. 5.10.

x 10 -1 0

Time
Fig. 5.12 Steady state error for 2 H z command, batch update every 5p.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
118

cos

sin

Time
Fig. 5.13 Steady state error to 1/5 Hz sine and cosine input, batch update every 5p.

2/5 Hz

4/5 Hz

Time
Fig. 5.14 Steady state error to 2/5 Hz and 4/5 Hz sine inputs, batch update every 5p.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
119

s_
O
H
M
C/2
C /2

U !
1 2 3 4 3
Time
Fig. 5.15 Steady state error to 8.2 Hz sine input, batch update every 5p.

\
0.5

LU 0
C/2
C /2

-0.5

-1
1 1-2 1.4 1.6 1.8
Time
Fig. 5.16 Steady state error to 20.2 Hz sine input, batch update every 5p.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
120

x 10 ’ 12

10p

2.5
coT
Fig. 5.17 Difference in frequency response amplitudes o f jS^I and \Sr\ for 5 and 10p, for
frequencies periodic with period p.

Co
1

2.5
coT
Fig. 5.18 Enlarge scale graph o f Fig. 5.17 for 10p.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
121

C hapter 6

C onclusions

Iterative learning control and repetitive control aim to improve the

performance o f tracking control systems by learning from experience to decrease the

error in present operation. We have developed understanding o f fundamental

properties o f linear iterative learning and repetitive control as summarized below.

Establishing stability o f a learning or repetitive control system is the first

design requirement, and stability conditions are discussed in chapter 2. Results show

that a frequency response based stability condition is a sufficient condition for

convergence to zero tracking error in both ILC and RC. In RC the advantages o f this

condition over the stability condition defining the actual boundary o f stability are the

ease o f com putation and the fact that the condition implies monotonic decay o f error.

In chapter 3 methods are developed to predict final error level o f ILC and RC

when using linear phase lead com pensator as well as zero-phase filter. These predict

the steady state error levels that will be reached for each choices made by the control

system designer. These methods can be used to optimize the choice o f the learning

process.

In chapter 4 an extension o f the Bode Integral Theorem is developed to apply

to learning control problems using the sensitivity function generated in chapter 3. The

Bode Integral Theorem is a fundamental constraint o f feedback control systems that

establishes the waterbed effect. The waterbed effect states that the transfer function

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
122

from command to the resulting error, or from output disturbance to resulting error,

has the property that attenuation in some frequency range m ust be paid for by

amplification in some other frequency range. We develop the Generalized Bode

Integral Theorem that shows how the waterbed effect is bypassed by learning control

law designs. This includes introducing more zeros than poles, modifying the gain in

the sensitivity transfer function, and introducing poles outside the unit circle with

non-causal filtering.

In chapter 5, the relationship between the waterbed effect and repetitive

control are studied using filtering for a frequency cut o ff the learning for the purposes

o f stabilization. By using filtering in repetitive control systems it becomes much

easier to satisfy the stability condition. It is shown that when using filtering in a real

time repetitive control mode, the frequency response stability condition is only

slightly more restrictive than the Nyquist stability condition defining the stability

boundary. This result agrees with similar results presented in chapter 2. In

engineering practice, one prefers to use the simple frequency response based stability

condition because o f the ease o f computation. It is shown that for real time repetitive

control, one is subject to the waterbed effect. On the other hand, it is possible to

bypass this effect for those frequencies having the period o f interest, by making batch

updates o f the repetitive control signal. Sufficiently slow batch updates in repetitive

control will result in the same situation that applies to iterative learning control for

these frequencies, where it is possible for certain classes o f learning control laws to

bypass the waterbed effect. Methods are developed to predict the disturbance to error

frequency response characteristics for the range o f possible batch updates rates.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
123

R eferrences

1. Arimoto, S., Kawamura, S., and M iyazaki, F., "Better operation o f robots by
learning." Journal o f Robotic System. 1984, V o l.l, p p .123-140

2. Bode, H.W., Netw ork Analysis and Feedback Am plifier Design. Princeton. NJ:
Van Nostrand, 1945.

3. N. K. Bose, D igital Filters, Theory and Applications, North-Holland, Amsterdam,


1985, p. 162.

4. Casalino, G. and Bartolini, G., "A learning procedure for the control movements
o f robotic m anipulators," IASTED Sym posium on Robotics and Automation.
Amsterdam, The Netherlands, 1984, pp. 108-111.

5. Craig, J. J., "Adaptive control o f manipulator through repeated trial," Proceedings


o f the Am erican Control Conference, San Diago, USA, 1984, pp. 1566-1573.

6. Elci, H., Longman, R.W ., Juang, J.-N., and Ugoletti, R., "Discrete frequency
based learning control for precision m otion control," Proceedings o f the 1994
IEEE International Conference on Systems, Man, and Cybernetics, San Antonio.
TX, USA, 1994, pp. 2767-2773.

7. Elci, H., Longman, R.W., Juang, J.-N., and Ugoletti, R., "Experiments in the use
o f learning control for maximum precision robot trajectory tracking," Proceedings
o f the 1994 Conference on Information Science and Systems, D epartm ent o f
Electrical Engineering , Princeton University, Princeton, NJ, USA, 1994, pp. 951 -
958.

8. Elci, H., Longman, R. W., Phan M. Q., Juang, J.-N. and Ugoletti, R . , "Automated
Learning Control throught Model Updating for Precision M otion Control."
Adaptive Structures and Composite M aterials: Analysis and Applications, AD-
Vol. 45/M D-Vol. 54, AS M E 1994, pp. 299-314.

9. Francis, B., and Zames, G., "On H-infinity optim al sensitivity theory for SISO
feedback system," IEEE Transactions on Autom atic Control, 1984, Vol. 29, no. 1.
pp. 9-16.

10. Hara, S., Omata, T., and Nakano, M., "Stability o f repetitive control systems."
Proceedings o f th 24th IEEE Conference on Decision and Control, Fort
Lauderdale, FL., 1985, pp. 326-327.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
124

11. Hara, S., and Yamamoto, Y., "Synthesis o f repetitive control system s and its
application, " Proceedings o f th 24th IEEE Conference on Decision and Control.
Fort Lauderdale, FL., 1985, pp. 1387-1392.

12. Hsin, Y.-P., Longman, R. W., Solcz, E. J., and de Jong, J., "Experimental
Com parisons o f Four Repetitive Control Algorithms," Proceedings o f th 31st
Annual Conference on Information Sciences and Systems. John Hopkins
University, Department of Electrical and Computer Engineering, Baltimore,
M aryland, 1997, pp.854-860.

13. Hsin, Y.-P., Longman, R. W., Solcz, E. J., and de Jong, J., "Experiments bridging
learning and repetitive control," Advances in the Astronautical Sciences, 1997,
Vol. 95, pp. 671-690.

14. Hsin, Y.-P., Longman, R. W., Solcz, E. J., and de Jong, J., "Stabilization due to
Finite W ord Length in Repetitive and Learning Control," Advances in the
Astronautical Sciences, Vol. 97, 1997, pp. 817-836.

15. Huang, Y.-C. and Longman, R. W ., "The Source o f the Often Observed Property
o f Initial Convergence Followed by Divergence in Learning and Repetitive
Control," Advances in the Astronautical Sciences, Vol. 90, 1996, pp. 555-572.

16. Inoue, T., Nakano, M., and Iwai,S., "High accuracy control o f a protron
cynchrotron m agnet power supply," Proceedings o f the 8th World Congress o f
IFAC, XX, 1981, pp. 216-221.

17. Lang, S., Complex Analysis, 3rd edition, Springer-Verlag, New York, 1993.

18. Longman, R. W. "Designing Iterative Learning and Repetitive Controllers,”


chapter in Iterative Learning Control: Analysis, Design, Integration and
Applications, S. Bien and J. Xu editors, Kluwer Academic Publishers. Boston.
1998, pp. 107-146.

19. Longman, R. W., "Iterative learning control and repetitive control for engineering
practice," International Journal o f Control, 2000, Vol. 73, no. 10, pp. 930-954.

20. Longman, R. W., "Real time HR zero-phase low-pass filtering for robust
repetitive control," Proceedings o f the 6th International Conference on Control,
Automation, Robotics and Vision, Singapore, December 2000, to appear.

21. Longman, R.W. and Songchon, T., "Trade-offs in designing learning/repetitive


controllers using zero-phase filtering for long term stabilization," Advances in the
Astronautical Sciences, 1999, Vol. 102, pp. 243-262.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
125

22. Longman, R. W. and Wirkander, S.-L., "Automated Tuning Concepts for


Iterative Learning and Repetitive Control Laws," Proceedings o f the 37th IEEE
Conference on Decision and Control, Tampa, Florida, USA, December 16-18,
1998, pp. 192-198.

23. M iddleton, R. H. , Goodwin, G. C. and Longman, R.W., "A method for


improving the dynamic accuracy o f a robot performing a repetitive task,"
University o f Newcastle, Newcastle, Australia, Department o f Electrical
Engineering Technical Report EE8546. Also , International journal o f Robotics
Research, 1989, Vol. 8, pp. 67-74.

24. Nakano, M., and Hara, S., "M icroprocessor-based repetitive control,"
Microprocessor-Based Control Systems, 1986, pp. 279-296.

25. Oh, S. J., Longman, R. W. and Phan, M. Q., "Use o f Decoupling Basis Functions
in Learning Control for Local Learning and Improved Transients," Advances in
the Astronautical Sciences, AAS/AIAA Spaceflight Mechanics Meeting, 1997,
pp. 651-670.

26. Omata, T., Nakano, M., and Inoue, T., "Application o f repetitive control method
to multivariable systems," Transactions o f SICE, 1984, Vol. 20, pp.795-800.

27. Phan, M. Q. and Longman, R.W., "A mathematical theory of learning control for
linear discrete multivariable systems," Proceedings o f the AIAA/AAS
Astrodynamics Conference, Minneapolis, Minnesota, USA, 1988, pp.740-746.

28. Pierre, D. A., "Reformulated Nyquist criterion for discrete-time systems," IEEE
Transactions on Education, 1989, Vol. 3, n o .l, pp. 59-61.

29. Plotnik, A. and Longman, R. W., "Subtleties in the use of zero-phase low-pass
filtering and cliff filtering in learning control," Proceedings o f the 1999
AIAA/AAS Astrodynamics Conference, Girdwood, Alaska, Advances in
Astronautical Science , Vol. 103, 1999, pp.673-692.

30. Seron, M. S., Braslavsky, J.H., and Goodwin, G. C., Fundamental Limitations in
Filtering and Control, Springer Verlag, 1997.

31. Songchon. T. and Longman, R.W., "Comparison o f the stability boundary and the
frequency response stability condition in learning and repetitive control," Second
International Workshop on Multidimensional (nD) Systems, Czocha Castle,
Lower Silesia, Poland, 2000, pp. 121-126.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
126

32. Songchon, T. and Longman, R. W., "Iterative learning control and w aterbed
effect," Proceedings o f the 2000 AIAA/AAS Astrodynamics Conference, Denver,
Cororado, 2000, pp. 444-453.

33. Songchon, T. and Longman, R. W., "Trade-offs in designing learning/repetitive


controllers using zero-phase filtering for long term stabilization," Advances in the
Astronautical Sciences, 2001, to appear.

34. Tomizuka, M., Tsao, T.-C., and Chew, K.-K.," Analysis and synthesis o f discrete
time repetitive controllers," Journal o f Dynamic Systems, M easurement, and
Control, 1989, Vol. 111, pp. 353-358.

35.U chiyam a, M., "Form ulation of high-speed m otion pattern o f a m echanical arm
by trial (in Japanese)," Transactions o f the Society fo r Instrumentation and
Control Engineer, 1978, Vol. 14, pp. 706-712.

36. Wang, Y. and Longm an, R. W., "Use o f Non-Causal Digital Signal Processing in
Learning and Repetitive Control," Advances in the Astronautical Sciences. Vol.
90, 1996, pp. 649-668.

37. Wen, H. P., Phan, M. Q. and Longman, R. W., "Bridging Learning Control and
Repetitive Control using Basis Functions," Advances in the Astronautical
Sciences, 1998 AAS/AIAA Spaceflight M echanics Meeting, Vol. 99. 1998,
pp.335-354.

38. Wen, H. P. and Longman, R. W., "Floquet Stability Analysis o f Disturbance


Cancellation using Iterative Basis Function Feedback," Advances in the
Astronautical Sciences, Vol. 102, 1999, pp.783-802.

39. Wirkander, S.-L. and Longman, R. W., "Lim it Cycles for Improved Perform ance
in Self-Tuning Learning Control." Advances in the Astronautical Sciences. Vol.
102, 1999, pp. 763-781.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

You might also like