You are on page 1of 16

94 Int. J. Mechatronics and Automation, Vol. 3, No.

2, 2013

Robust control stability using the error loop

Enrico Canuto*, Wilber Acuña-Bravo and


Carlos Perez Montenegro
Politecnico di Torino,
Department of Control and Computer Engineering,
Corso Duca degli Abruzzi 24, 10129 Turin, Italy
E-mail: enrico.canuto@polito.it
E-mail: wilber.acunabravo@polito.it
E-mail: carlos.perez@polito.it
*Corresponding author

Abstract: This paper formulates the error loop as a tool for designing robust stability control
systems in the presence of structured and unstructured uncertainties. The error loop indicates that
uncertainties can be accommodated through the design of the noise estimator, which is the unique
feedback channel from plant to control. The real-time model that is embedded in the control unit
and the noise estimator constitute a state predictor. The embedded model consists of a
controllable dynamics plus a disturbance dynamics fed by the noise estimator. It is shown that
causality constraint prevents perfect cancellation of causal uncertainties (unknown disturbance),
but makes the control law which is fed by the state predictor to play a role, thus offering a further
degree of freedom. Employing asymptotic expansions of the closed-loop transfer functions,
simple, explicit design formulae derive from stability inequalities. They relate closed-loop
eigenvalues to model parameters and requirements, and define an admissible frequency band for
the state predictor bandwidth. This paper restricts formulation to the univariate case. A simple
example is provided with simulated and experimental data.

Keywords: robust control; closed-loop; stability; error loop; embedded model control.

Reference to this paper should be made as follows: Canuto, E., Acuña-Bravo, W. and
Perez Montenegro, C. (2013) ‘Robust control stability using the error loop’, Int. J. Mechatronics
and Automation, Vol. 3, No. 2, pp.94–109.

Biographical notes: Enrico Canuto received his degree in Electrical Engineering from
Politecnico di Torino, Turin, Italy, where he joined the staff as an Associate Professor of
Automatic Control in 1983. From 1982 to 1997, he contributed to data reduction of the European
astrometric mission Hipparcos. Technological studies in view of scientific and drag-free space
missions, like Gaia and GOCE, provided the opportunity of applying embedded model control to
drag-free satellites and to electro-optics. He contributed to the conception, design and
implementation of the nanobalance interferometric thrust-stand, capable of sub-micronewton
accuracy. Presently, he is involved in the design of the orbit, formation and attitude control of the
Next Generation Gravity Missions of the European Space Agency. His research interests cover
all the entire field of control problems that are challenging because of complexity, uncertainty
and precision.

Wilber Acuña-Bravo received his degree in Electrical Engineering from Universidad del
Quindío, Armenia, Colombia, and MS in Control Engineering from Universidad de Los Andes,
Mérida, Venezuela. In 2011, he received his PhD in Information and System Engineering from
Politecnico di Torino, Turin, Italy, where he is currently a Research Assistant. His research
interests include electro-hydraulic control systems and control applications.

Carlos Perez Montenegro received his degree in Electrical Engineering from Pontificia
Universidad Javeriana, Bogotá, Colombia. Currently, he is a PhD student in Information and
System Engineering at Politecnico di Torino, Turin, Italy. His main research interests include
space control applications and quadricopter design and control.

Copyright © 2013 Inderscience Enterprises Ltd.


Robust control stability using the error loop 95

1 Introduction since model-plant discrepancies encoded in e are attenuated.


The difficulty is that h, being output and command
1.1. Uncertainty and the error loop dependent cannot belong to Du, as the latter is defined as a
Closed-loop performance is affected by uncertain class of command independent signals. This is one of the
discrepancies between model and reality. Three kinds of chief instability sources of any feedback control design,
discrepancies may be distinguished (Maciejowski, 1989; since feedback signals – in this case the estimated
Doyle et al., 1992) as in Figure 1. innovations wd and wu – spill components incoherent with
Causal uncertainty, denoted by clouds, regards the design model. Appropriate stability conditions need to
unpredictable actions, referred to as innovation, and be formulated and proved, which are shown to be mainly
formulated either as a white-noise or a bounded and related to the noise estimator (and therefore to the state
arbitrary-signal class corresponding to wd and wu in predictor) design. Specifically, frequency-domain design
Figure 1. They are independent of the model state and suggests to enlarge the state predictor bandwidth (BW) to
command and only affect performance. The mechanism allow estimation and cancellation of the dominant low
through which they affect a plant is not merely frequency components.
unpredictable, as they accumulate prior to be released. Unstructured uncertainty, denoted by a cloudy 3D box,
Causes of this kind are referred to as ‘unknown’ disturbance refers to model-class discrepancies and uncertainties, which
and their model, the disturbance dynamics D in Figure 1, are unavoidable due to the embedded model finite order.
generates the signal class du ∈ Du. The control goal is to They may take several forms as in Maciejowski (1989),
cancel their effect less the innovation, as the latter occurs Doyle et al. (1992) and Ross Barmish (1994). The output
while the command is acting on the plant. Innovations wd fractional form [multiplicative form in Doyle et al. (1992)]
and wu are necessary and sufficient for updating the state of is adopted here as in Canuto (2007) and in Canuto et al.
D. They must be real-time estimated by correlating them to (2012). They are driven by the model output ym and yield
the model error e = y – ym through a noise estimator as in the model error e (plant minus model) as output, as follows:
Canuto et al. (2010). This is the basic mechanism of the
e = y − ym = ∂P ( ym ) + wy , (1)
‘disturbance observers’ pioneered by Johnson (1971), and
Hostetter and Meditch (1973), and further studied and where y is the measured plant output. The dynamic operator
applied by Mita et al. (1998), Bickel and Tomizuka (1999), ∂P is the (uncertain) ‘neglected dynamics’, a term
Schrijver and van Dijk (2002), Choi et al. (2003), and preferable to unstructured uncertainty, and wy is the
Katsura et al. (2007). The embedded model control of measurement error, which has been assumed adding to ∂P.
Canuto (2007) inspired by Donati and Vallauri (1984) Components of (1) are completely neglected in the
develops around the concept that noise is the necessary and embedded model. Spilling e to the command through
sufficient feedback from plant to control. The ensemble of feedback must be restricted, since, for what concerns ∂P, it
noise estimator and design model, to be embedded as a can be assimilated to a command-dependent output error. In
real-time model in the control unit – hence the name addition, wy prevents accurate estimation of the innovations
embedded model – implements a one-step state predictor. wd and wu, which difficulty leads to Kalman filter
The structured (or parametric) uncertainty h(·), denoted optimisation (see Kwakernaak and Sivan, 1972). The
by a cloudy 3D box, either refers to discrepancies from the uncertainty-based design restricts noise estimation in the
model class in the form of parameter uncertainty or to frequency band (usually at higher frequencies below
neglected relations between model variables (cross Nyquist frequency) where ∂P dominates.
couplings). They must be distinguished from the known As a result in robust design, structured and unstructured
interconnections, that are denoted with h (⋅) and must be uncertainties must be accommodated by a trade-off in the
treated as known disturbances. They have been extensively design of the state predictor BW: a wider band is required to
studied in the literature, as for instance by Chapellat et al. cancel structured discrepancies, a narrower band to prevent
(1990), Foo and Soh (1993), Ross Barmish (1994), spilling of neglected dynamics. The design defines a
Calafiore and Dabbene (2002), Chen et al. (2005), stability interval where to place noise-estimator gains or,
Patre et al. (2008). They may take several forms, equivalently, the state-predictor eigenvalues. Converting
from command-to-state to output-to-state relations: only eigenvalues into frequency domain, lower and upper bounds
state-to-command relations are treated here since they affect of the state-predictor BW are obtained. The interval width
the model eigenvalues. They are included in the so-called depends on the ratio of the predictor and control-law
design model and implemented as a numerical simulator, eigenvalues, and on a stability margin. When the interval is
but they are forcedly neglected in the embedded model, void, robust design becomes unfeasible, which may be due
except for the known part h, and are surrogated by the to excessive stability margin (conservative design) or large
unknown disturbance class Du. The advantage is that h can uncertainty (poor modelling).
be made implicitly known by du and therefore cancelled by
u, thus favouring robustness of the model-based design
96 E. Canuto et al.

Figure 1 Block-diagram of embedded model and uncertainties (see online version for colours)

Figure 2 Block-diagram of the error loop (see online version for colours)

wu

Disturbance du
wd dynamics
D

Structured
uncertainty
h ( ⋅)
Control unit e
h (⋅) Dynamics d y State
Known ey
predictor e −
disturbance
M Sm
eˆy
wy

y Neglected e State eˆ y ey
− −
dynamics ∂P predictor
− ym
Vm Causality
S c ML w
Uncertainties

Stability inequalities are obtained in Section 3 from error e defined in (1), the tracking error ey and the prediction
equations driven by model uncertainties in Figure 1. They error eˆ y defined by
generate a loop passing through different errors, and
therefore designated as the ‘error loop’ in Canuto (2007). ey = y − ym
Three main errors enter the loop in Figure 2: the model error (2)
eˆ y = ym − yˆ m ,
Robust control stability using the error loop 97

where y is the reference signal driving the loop and yˆ m is 1 the modelling procedure pivoting on the disturbance
the one-step prediction of the model output ym. None of the dynamics D and on the driving noise as the sole
above errors is measurable as they include the model output plant-to-model feedback
ym, which is only available either as a signal class or from 2 the performance equation (63) including all the
simulated runs. The corresponding measurable errors that uncertainty contributions and their closed-loop
are indicated as the output of the control unit in Figure 2, accommodating filters
are made available by replacing ym either with yˆ m or y in
3 the derivation of (63) in Section 3
(2). They correspond to the measured model error e , to the
4 the asymptotic inequalities in Section 3 and their
measured tracking error eˆy and to the control error ey. They
derivation throughout this paper.
are defined as follows
More sophisticated algorithms, simulation and in-field
e = y − yˆ m = e + eˆ y tuning may refine the resulting robust design.
eˆy = y − yˆ m = ey + eˆy (3) Formulation employs discrete-time transfer functions
denoted as y(z) = M(z)u(z). Sometimes z is dropped. Since
e y = y − y = eˆy − e = ey − e.
this paper aims to be introductory, the univariate case is
treated.
1.2 Content of this paper
In Section 3, this paper concentrates on the derivation and 2 The embedded model and the uncertainty
properties of the error loop. The main result is the
following: structured and unstructured uncertainties may be 2.1 The embedded model and its realisation
accommodated by the state predictor sensitivity Sm and its
2.1.1 Generalities
complement Vm as shown in Figure 2. The result seems
partly departing from most of the literature, where either the A discrete-time dynamics is considered, which is associated
control law is dedicated to the purpose as in Solihin et al. to the time unit T, to the Nyquist frequency fmax = 0.5/T and
(2011) or no distinction between control and state predictor to discrete times iT. The model output ym is the response to
is made. On the contrary, structured uncertainties must be the command u and to the disturbance d as follows (see
real-time estimated so as to update the embedded model. Figure 1):
This is obtained by shaping the lower-frequency part of the
ym ( z ) = M ( z ) ( u ( z ) + d ( z ) ) = M ( z )u ( z ) + d y ( z )
sensitivity Sm. Unstructured uncertainties must be blocked (4)
from spilling into the embedded model in order to prevent d ( z ) = du ( z ) + wu ( z ) + h ( xc ) + h ( x c ) ,
instability. This is achieved by shaping the higher-frequency
part of Vm. Actually, as shown in Section 3, because of where
causality, which prevents cancelling innovation wu in
M = Cc ( zI − Ac ) Bc = Cc N ( z ) / φ ( z ).
−1
(5)
Figure 1, – wu and wd can only be causally estimated –, the
state predictor is affected by the correction Sw = ScMLwSm is irreducible and of order nc. To be simple, CcN(z) is
in Figure 2, right bottom, which depends on the control-law assumed to be Hurwitz. The input signals u and d are scaled
sensitivity Sc, on the controllable dynamics M and on the to the output units; dy is the output disturbance to be used in
noise estimator Lw. Causality adds a degree of freedom and Section 3. The disturbance d is decomposed into the
shows how state predictor and control law intertwine in the following components:
overall control sensitivity S and its complement V.
Shaping Sm and S is done by approximating them and 1 The unknown disturbance du(z) = D(z)wd(z) is driven by
their complements Vm and V with low- and high-frequency a noise vector wd of size nw through the irreducible
asymptotes, respectively, within the Nyquist frequency dynamics D = Mc(zI – Ad)–1Gd of order nd, with state
band. Asymptotes can be related to noise estimator and vector xd. The disturbance must be rejected by the
control gains and to state predictor and state feedback control action.
eigenvalues. The intersections of the asymptotes with 2 The unpredictable noise wu(z) cannot be cancelled due
the zero dB line are interpreted as the bandwidths of to causality, since any sample wu(i) occurs while the
(asymptotic) sensitivity and complement. The main result command sample u(i) actuates plant and model.
affirms that a robust design accommodating structured and
unstructured uncertainty defines an admissible region Δfm of 3 The non-linear feedback h(xc ) + h (xc ) depends on the
the BW fm of the state predictor sensitivity Sm. The state xc of M; the unknown part h(xc) is treated as a
admissible region in turn depends on the control ratio γ = fm structured uncertainty to be estimated and rejected by
/ fc, fc being the BW of Sc, and on the stability margin η–1 > the control action.
1. Machinery and proof of the asymptotes are carried out
step by step throughout the paper. The disturbance dynamics D plays the role of the weighting
Main contributions are credited to be: functions in Maciejowski (1989) and Doyle et al. (1992),
but is designed to be parameter-free and explicitly coded in
98 E. Canuto et al.

the control unit. A disturbance adding to the command is the neglected dynamics, as well as amplifier dynamics and
simplest case, which is referred to as the collocated case. LuGre friction dynamics (Canudas de Wit et al., 1995).
The not collocated case is treated in Canuto (2007). Signals Transfer functions and h(·)in (4) hold
wd and wu may be treated as white noise, but when statistics
is unknown they may be also interpreted as bounded, ⎡ ym = θ ⎤
M ( z ) = ( z − 1) −2 , xc = ⎢ ⎥
arbitrary signals, whose average converges to a null value. ⎣ Δθ ⎦
Null value is therefore their best prediction. ⎡ wa ⎤ ⎡a ⎤
All the state variables of M and D are observable by the D( z )w d ( z ) = ( z − 1) −1 ⎣⎡1 ( z − 1) −1 ⎦⎤ ⎢ ⎥ ( z ), xc = ⎢ ⎥ (9)
w
⎣ s⎦ ⎣s⎦
model output ym. Low- and high-frequency asymptotes of M
for f < fmax are as follows: h ( xc ) = − ( sgn(Δθ ) A (| Δθ |) − mgl cos ( ym ) ) T 2 / J

lim f →0 M ( jf ) = ( f 0 / f )
m0 h ( xc ) = 0.
(6)
lim f →∞< fmax M ( jf ) = ( f ∞ / f )
m∞
, Parameters in (6) can be proved to hold
m0 = m∞ = 2, d 0 = d ∞ = 4
where m0 > 0 and m∞ > 0 imply a pole-zero excess of at (10)
least one. Since D is sized 1 × nw, it is post-multiplied times f 0 = f ∞ = f d 0 = f d ∞ = (2πT ) −1.
the arbitrary transfer vector Ld(z), nw × 1, in a way that no
The block diagram with input, state and output variables is
zero-pole cancellation occurs, and
in Figure 3. The noise size is nw = 3 and the embedded
lim z →1 L d ( z ) = Ld 0 < ∞ model size parameters are n = d0 = d∞ = 4. The
(7)
lim z →∞ L d ( z ) = Ld ∞ < ∞. block-diagram includes the reference dynamics to be
explained in Section 2.3. Boxes marked with Σ represent
Then, given Ld(z) satisfying (7), we assume discrete-time integrators having the generic state equation
lim f →0 M ( jf )D( jf )L d ( jf ) = ( f d 0 / f )
d0
x(i + 1) = x(i ) + u (i ), x(0) = x0 . (11)
(8)
lim f →∞< fmax M ( jf )D( jf )L d ( jf ) = ( f d ∞ / f )
d∞
, Initial states are marked by a superimposed arrow.

where d0 – m0 > 0 and d∞ – m∞ > 0 imply that DLd possesses Figure 3 Block diagram of the embedded model of the example
at least a unitary pole. The lim f →∞< fmax accounts for (see online version for colours)

asymptotes in the region f < fmax: they are due to poles in


excess of the type (z – 1 + β)–1 with 0 ≤ β < 1, and thus [ rad ] [ rad ] y (i )
[ rad ]
Δθ ( i ) e (i )
exclude pure delays. u (i ) ym ( i ) = θ ( i ) −
∑ ∑
The noise wd is multivariate, since D is the composition
of different stochastic signals, and each component is driven
by different noise components. Moreover, M and D are [ rad ] [ rad ]
assumed to be strictly causal, which implies that the relevant a (i ) s (i ) ws ( i )
∑ ∑
output ym and du are linear combinations of state vectors wa ( i )
only. D is usually stylised from experimental spectral wu ( i )
densities by whitening their perturbing sources as in Canuto
and Rolino (2004), and Canuto (2008). The embedded
model control implements the control unit around equation
Δθ ( i )
(4) written in the form of a state equation. This entails the u (i ) au ( i ) y (i )
∑ ∑
need of estimating the noise vectors wd(i) and wu(i) for
updating controllable and disturbance states. As in Kalman
filtering (see Kwakernaak and Sivan, 1972), the only way to Similar design model and considerations apply to a ball and
the purpose is to extract the causal estimates w and wu beam device (Keshmiri et al., 2012), when ball position is
from the model error e defined in (1). uncontrolled. The ball control would require a more
complex dynamics than (9).
2.1.2 Example
2.2 Noise estimation and output prediction
Consider a balanced robot arm moving, in a vertical plane, a
mass m that is distant l from the rotation centre. The 2.2.1 Generalities
counter-clockwise rotation θ with respect to the horizontal
Noise estimation forces the embedded model to update and
plane is denoted with ym. The angular rotation ω = y m is instantiate the signals ym(i) and du(i) into the realisations
replaced by the angular increment Δθ = ωT. The arm is yˆ m (i ) and dˆu (i ). The mark ^ denotes one-step prediction as
driven by a DC motor and is subject to gravity torque and
friction. The total inertia on the gear output shaft is J. Gear the relevant signals only depend on the past measures
backlash and torsional deformations are confined into y(i – k), k > 0. On the contrary, the noise estimates w d (i )
Robust control stability using the error loop 99

and wu , and the actual (or measured) model error embedded model are one-step anticipated as indicated by
e (i ) = y (i ) − yˆ m (i ) are barred, meaning they depend on the the boxed z.
present measure y(i). Thus, one must distinguish between
Figure 4 Embedded model and noise estimator (see online
the a priori embedded model (4) and the a posteriori model version for colours)
which is implemented in the control unit and is forced by a
realisation e of e. Consequently, unknown prediction errors from plant
y (i )
establish between model signals and their prediction as
follows: u (i ) Controllable yˆ m ( i ) e (i )
dynamics
eˆ y = ym − yˆ m d̂ ( i ) M
(12)
eˆd = du + h ( xc ) − dˆu . z
Known cross Noise
coupling estimator
Defining the noise error as ew = wu − wu , a first expression h ( ⋅)
relating ê to the errors in (12) is z
wd (i )
dˆu ( i ) Disturbance
eˆ y ( z ) = M ( z ) ( eˆd + ew ) ( z ). (13) dynamics
Ld
D
To prove it with the help of (4), rewrite the former equation z wu ( i )
in (12) as follows: Lw
Embedded model
ym − yˆ m =M ( u + du + wu + h ( xc ) + h ( xc ) )
(14)
(
−M u + dˆu + wu + h ( xc ) . ) The coefficients of Ld and Lw, i.e., the noise estimator gains,
are fixed by the state-predictor closed-loop eigenvalues, i.e.,
Equation (14) yields (13) with the help of the second
by the poles of Sm. The closed-loop spectrum is denoted as:
equation in (12).
A further expression derives from the linear and Λ m = {1 − γm1 , ..., 1 − γmk , ..., 1 − γmμ } , (20)
time-invariant noise-estimator, which is driven by the
measured model error e and is partitioned as follows: where μ ≥ n = nc + nd is the order of the state predictor.
Eigenvalues in (20) are written in terms of the
w d ( z ) = L d ( z )e ( z ) complementary eigenvalue γmk. When the latter is real and
(15)
wu ( z ) = L w ( z )e ( z ), 0 < γmk ≤ 1, provides the frequency fmk [Hz] through
γmk = κm2πfmkT, where π–1 ≤ κm ≤ 1 and κm → 1 as γmk → 0.
where Ld and Lw are suitable transfer functions having finite
Asymptotic inequalities in Section 3 will be obtained by
gains for z → 1 and z → ∞, i.e.,
assuming equal and real eigenvalues, i.e.,
lim z →1 L w ( z ) = lw0
(16) γmk = γm = 2πf mT , k = 1, ..., μ, (21)
lim z →∞ L w ( z ) = lw∞ .
having set κm = 1, i.e., γm << 1. The frequency fm is referred
Specifically Lw must be all-pass for estimating the white to as the state-predictor frequency.
noise wu, whereas Ld may be low-pass. Replacing (15) in The low-frequency asymptote of Sm follows from (8)
(13) and dropping z yields and holds
eˆ = M ( h ( xc ) + du + wu + h ( xc ) ) − M ( DL d + L w ) e S m 0 ( jf ) = lim f →0 S m ( jf )
(17)
= d y − MHeˆ − MHe,
= lim f →0 (1 + M ( DL d + L w ) )
−1

where the output disturbance dy in (4) has been used


1 (22)
together with the overall feedback H = DLd + Lw and the = lim f →0
1 + ( f d 0 / f ) + lw 0 ( f 0 / f )
d0 m0
key error equality
d0
e = e + eˆ y . (18) ⎛ f ⎞
=⎜ ⎟ ,
Assumptions on Ld and Lw imply H to be proper. Inserting ⎝ f m0 ⎠
the sensitivity Sm = (1 + MH)–1 and the complement where fm0 = fd0. The high-frequency asymptote of Vm
Vm = 1 – Sm into (17), yields the second expression of eˆ, follows from (6) and holds
namely
lim f →∞< fmax Vm ( jf ) = lim f →∞< fmax M ( DL d + L w )
eˆ y ( z ) = − Vm ( z )e( z ) + S m ( z )d y ( z ), (19) m∞
⎛ f ⎞
which is illustrated in Figure 2. = ⎜ m∞ ⎟ (23)
⎝ f ⎠
Figure 4 shows the noise estimator driving the
f m∞ = f ∞ ( lw∞ )
1/ m∞
disturbance dynamics. All the output variables of the .
100 E. Canuto et al.

The frequencies fm0 and fm∞ are referred to as the sensitivity φm ( z ) = ( z − 1)4 + l0 ( z − 1)2 + l1 ( z − 1) + l2 . (27)
and complement BWs, and are proportional to fm through
the coefficients κm0 and κm∞ that depend on the spectrum Λm Since the polynomial in (27) lacks the third degree
in (20). They hold coefficient, no set {l0, l1, l2} exists capable of stabilising Sm.
The only alternatives are either to add a fictitious noise
f m 0 = κm 0 f m on the angular increment Δθ in Figure 3, as it is done by
f m∞ = κm∞ f m (24) Kalman filters, or to employ a dynamic estimator capable of
lw 0 = κ w0 f mnc , respecting noise layout and size. As an alternative
interpretation of the dynamic estimator, we may think of
where the polynomial relation of lw0 in (16) has been added. replacing the missing input noise with a further output q
which is linearly independent of e . In fact, any dynamic
2.2.2 Example filter, in the limit a delay, makes output independent of
input. As a result, the number of available feedback
Since the noise size nw = 3 is less than the order n = d0 = channels from model error to noise double, from three to
d∞ = 4 of the noise-to-output transfer function MD, a six. It can be shown that stability can only be recovered by
dynamic noise estimator is mandatory for stabilising the tuning the six gains together with the parameters of the
closed-loop state predictor (embedded model and noise dynamic filter from the error e to the output q.
estimator), as in Canuto et al. (2010). To prove it, assume Assume a first-order filter q = ( z − 1 + β ) −1 e , because
first a static noise estimator in (15), namely:
the noise deficiency is just ne = n – nw = 1. The following
⎡ wa ⎤ ⎡la ⎤ noise estimator replaces (25):
⎢ w ⎥ ( z) = ⎢l ⎥ e ( z) (25)
⎣ a⎦ ⎣ s⎦ ⎡la ⎤ ⎡ ma ⎤
wu ( z ) = lu e ( z ). L d ( z ) = ⎢ ⎥ + ⎢ ⎥ ( z − 1 + β ) −1
⎣ ls ⎦ ⎣ ms ⎦ (28)
The corresponding loop transfer function MH = M(DLd + L w ( z ) = lu + ( z − 1 + β ) −1 mu .
Lw) in (17) is found to be
l0 ( z − 1) 2 + l1 ( z − 1) + l2
M ( z )H( z ) = , (26)
( z − 1) 4

and the denominator of Sm in (19) holds

Figure 5 Block diagram of the embedded model plus noise estimator and control law (see online version for colours)
Robust control stability using the error loop 101

The seven gains in (28) must be explicitly related to the 2.3 Reference dynamics and tracking error
closed-loop eigenvalues, or, that is the same, to the poles of
Performance refers to a class of reference signals y which
Sm, whose cardinality is μ = n + ne = 5. Equal eigenvalues
are assumed, namely Λm = {1 – γm, …, 1 – γm}. Since the is assumed to satisfy the same dynamics as in (4), but
seven gains in (28) satisfy the following five equations corrupted by a known disturbance du as follows:
β = 5γm y ( z ) = M ( z ) ( u ( z ) + du ( z ) ) . (31)
lu = lw0 = 10γm2 = κ w0 f m2
Here, du is assumed to be known, which is not the case
lu β + mu + la = 10γm3 (29)
when it is recorded from measurements. The reference
la β + ma + ls = 5γm4 signal class is shaped by the open-loop command u , which
ms = γm5 , is real-time computed by a reference generator (not treated
their selection may be optimised. Equation (29) shows that here, see Figure 5) capable of matching operator requests
the filter gain β is essential for recovering stability, as and technology limits. Typically, model predictive control
already mentioned. Moreover, either the pair (ma, mu) or (Camacho and Bordons, 2003) is concerned with it. The
reference state is denoted with x.
(la, ls) can be set to zero. The latter solution, i.e., la = ls = 0,
is preferable, since forcing them to be zero reduces the Control performance is assessed by means of the
contribution of e which is noisier than q. Using (29) and tracking error ey = y − ym , satisfying
(24), the asymptote parameters in (22) and (23) become
ey ( z ) = M ( z ) ( u + du − u − du − h(⋅) − wu − h (⋅) ) ( z )
(32)
( β / ms )1/ 4
γ ey ( z ) = Cc ( x − xc ) ( z ) = Cc e ( z ).
d 0 = 4, f m 0 = = 4 5 m = κm 0 f m
2πT 2πT (30)
In the example, du = 0, and the reference block-diagram is
l 10γm
m∞ = 2, f v∞ = u = = κ m∞ f m . shown in Figure 3, bottom. The reference state vector is
2πT 2πT
denoted by
Figure 5 shows the block-diagram of the embedded model
in Figure 3 plus the dynamic noise estimator (28) and the ⎡y = θ⎤
x=⎢ ⎥. (33)
control law to be outlined in Section 3.1.2. Gains are ⎣ Δθ ⎦
denoted with circles. Dashed circles correspond to zero
gains.
2.4 Model errors and uncertainties
Figure 6 Bode plots (magnitude) of the sensitivity transfer 2.4.1 Generalities
function and of the complement (see online version
for colours) Following the literature (Maciejowski, 1989), two kinds of
2 model errors are treated in addition to innovations wd and
10
wu.
Structured uncertainty is defined by the uncertain
0
‘static’ feedback h(xc) connecting xc to u as in (4). Being
10 unknown, it cannot be explicitly cancelled by the control
law in Section 3. The inclusion in (4) of the disturbance du
Magnitude

allows to surrogate h(·), but at the cost of closed-loop


-2
10 degradation since a feedback link is replaced by a
command-independent signal as du. Closed-loop stability
Complement V, γ=0.4, f m=1.9 Hz must be guaranteed. h(·) is assumed differentially bounded.
10
-4 The bound is obtained expanding h around the reference
Sensitivity S
state x as follows:
Predictor complement Vm

-6
Predictor sensitivity Sm h ( xc ) = h ( x ) + Δh ( x )( x − xc )
10 -1 (34)
10 10
0
10
1
10
2
x = x − α ( x − xc ) , 0 ≤ α < 1.
Frequency [Hz]
If x and the state tracking error e = x − xc are bounded,
Figure 6 shows typical Bode plots (magnitude) of the also x is bounded. Using (5) and (32), the first order
predictor transfer functions (dashed lines), and of the overall expansion in (34) is written as
transfer functions to be derived in Section 3.
102 E. Canuto et al.

Δh ( x ) N ( z ) The frequency square in (41) is imposed by nc = degφ = 2. fg


Δh ( x )( x − xc ) ( z ) = e ( z ), (35) is the ‘pendulum’ natural frequency of the arm, whereas
Cc N ( z )
2πf Δθ (Δθ ) is the friction/inertia pole, depending on the
and the parametric error eΔh is defined by reference angular rate Δθ , and such that f Δθ (Δθ ) ≤ f Δθ (0).
ΔH ( x , z ) Neglected dynamics is restricted to gear backlash and
eΔh ( z ) = M ( z )Δh ( x )( x − xc ) ( z ) = e ( z ), (36) torsional deformation. Due to backlash, dynamics is
φ ( z) non-linear. It can be linearly approximated for |e| ≥ emin, the
lower bound corresponding to the backlash width. Then
where φ has been defined in (5). ΔH ( x , z ) = Δh ( x ) N ( z ) is
∂P ( z )  − ( z − 1) ( ( z − 1) 2 + β t ( z − 1) + α t )
2 −1
a polynomial having degree mh = degΔH < degφ, and
bounded coefficients (42)
αt  ( 2πft T ) , βt  α t (α t + 2ζ t ) , pT = [ ft ζ t ] ∈ Π,
2

Δhk ( x ) ≤ Δhk ,max ( x ) , k = 0, ..., mh − 1. (37)


expresses the fractional error between the motor-gear-arm
The same polynomial having the bounds in (37) as dynamics having flexible transmission and the rigid body
coefficients is denoted with ΔH max ( x, z ). (9). The parameters in p are the uncertain natural frequency
Unstructured uncertainty is defined as an uncertain ft and the damping coefficient ζt. Typical Bode diagrams,
including a ball and beam device, are in Canuto et al.
dynamic operator ∂P(ym, …) from the model output ym to
(2012). The response peak and the peak frequency in (39)
the plant output y. Input-output stability and linear,
hold
time-invariance are assumed, implying that the model error
(1) has the expression f ∂P = ft ,min = min ft ∈Π ft
(43)
max p∈Π ∂Pmax ( z ) = ( 2ζ t ,min ) .
−1
e( z ) = ∂P ( z; p) ym ( z ) + wy ( z ), ∂P ( z; p) ∈ ∂P. (38)

where ∂P is the uncertainty class defined by a parameter


vector p ∈ Π belonging to a bounded set Π, and wy is the
measurement noise. The transfer function ∂P results from 3 Control law and the error loop
the sampled-data transform of a continuous-time dynamics 3.1 Control law and tracking error
enclosed between the plant zero-order hold and the output
sampler. Since ∂P is dropped from the embedded model, it 3.1.1 Generalities
may affect the overall stability, and calls for robust stability Given (4) and (31), in order to respect causality, the control
conditions. Here, we assume that there exists a worst-case law must depend either on the state variables or on their
element ∂Pmax(z; pmax) in ∂P such that combinations like ym and du, but not on wu. In addition, the
max f < fmax ∂Pmax ( f ) = max ∂P max f < fmax ∂P ( f ) control law must guarantee that the tracking error
ey = y − ym is bounded, and that the mean value tends
0 < arg max f < fmax ∂Pmax ( f ) = f ∂P (39)
asymptotically to zero. The control law
= min ∂P arg max f < fmax ∂P ( f ) < f max
u ( z ) − u ( z ) − du ( z ) = C( z ) ( e ( z ) + eˆ( z ) )
i.e., the peak of the worst case is the highest peak in the (44)
−du ( z ) − h ( xc ) + eˆd ( z )
class ∂P, and the peak frequency f∂P is the lowest in the class
∂P. can be shown to bound ey under appropriate conditions on
C, and on the prediction errors eˆ y and eˆd . In (44), C(z) is
2.4.2 Example usually improper with a bounded low-frequency gain
Using (9), the transfer function in (36) is found to be c0 = limz→1C(z) ≠ 0 and

ΔH ( x , z )
=
( )
Δh1 Δθ ( z − 1) + Δh0 θ ( ), m = 1, nc = 2 (40)
lim f →∞< fmax C( jf ) = c∞ f m∞ −1 . (45)
h
φ ( z) ( z − 1) 2 Inserting (44) into (32), simple manipulations yield
where the bounds of the dimensionless coefficients in (40) ( I + M ( z )C( z ) ) ey ( z ) = −M ( z )C( z )eˆy ( z )
are given in terms of the frequencies fg and f Δθ (Δθ ), i.e., (46)
−M ( z ) ( eˆd + ew ) ( z ) − M ( z ) wu ( z ).

( )
Δh1 Δθ Then, by replacing (13) in (46), the tracking error equation
follows:
= sgn(Δθ )dA ( Δθ ) / d | Δθ | T / J ≤ 2πf v ( Δθ ) T (41)
Δθ =Δθ
ey ( z ) = −eˆ y ( z ) − S c ( z )M ( z ) wu ( z ). (47)
( ) ( )
Δh0 θ = mglT 2 sin θ / J ≤ ( 2πf g T ) .
2

In (47), the control sensitivity Sc = (1 + MC)–1 makes its


appearance, and the sign of eˆ y is due to a different sign of
Robust control stability using the error loop 103

ym in the errors. The asymptotes of ScM and Vc = 1 – Sc can Performance only depends on the non-rejected noise wu
be found to be and initial conditions.

lim f →0 S c ( jf )M ( jf ) = lim f →0 (1 + MC) −1 M 2 Anti-causal law: it corresponds to include wu in (44),


which simplifies (47) to the ideal equality
( f 0 / f )m 0
1
= lim f →0 = ey = −eˆ y . (52)
1 + ( f0 / f )
m 0
c0 c0
(48)
Vc∞ ( jf ) = lim f →∞< fmax Vc ( jf ) = lim f →∞< fmax MC Moreover, adding eˆd = 0 and ew = 0 (model-based
assumptions), tracking error becomes zero, less the
c∞ ( f ∞ )
m∞
f c∞ free-response. Anti-causal laws cannot be realised since
= = .
f f u(i) depends on y(i – 1), whereas wu (i ) depends on
y(i). In terms of transfer functions, it would mean
The coefficients of C derive from the poles of Sc, are of the
Sc → 0 for f < fmax (the open switch in Figure 2 below,
order κ = nc ≥ m∞, and are designed to stabilise Sc. Likewise
ScMLw), which is unrealisable because of the Bode’s
in (20), the closed-loop spectrum is denoted with
integral theorem (Maciejowski, 1989; Wu and
Λ c = {1 − γc1 , ..., 1 − γck , ..., 1 − γcκ } , (49) Jonckheere, 1992).

and equal eigenvalues are assumed, γck = γc = 2πfcT, Clearly, (47) is a combination of the ideal equalities (51)
k = 1, …, κ. The control-law frequency is fc. Likewise in and (52).
(24) we can write Figure 7 shows the whole control unit that is built
around the embedded model: the control law interfaces
f c ∞ = κ c∞ f c model and plant, the noise estimator interfaces plant and
(50)
c0 = κc 0 f cnc model, and the reference generator interfaces operator/plant
and model. The delay z–1, from control law to plant,
Equation (47) may be better understood if one restricts to a balances the model one-step prediction. Digitisation
pair of ideal conditions: converts the computed command to a digital signal, and
may include a non-linear inversion. The opposite is denoted
1 Model-based control law: it corresponds to eˆd = 0,
with D/A and provides the embedded model with the same
ew = 0, to eˆ y = 0 in (13), and to the ideal equality in command dispatched to the plant.
(47)
ey = −S c Mwu . (51)

Figure 7 The control unit built around the embedded model (see online version for colours)
to plant from plant
y (i )

u (i ) Controllable yˆ m ( i ) e (i )

Digitisation z −1 D/A dynamics
d̂ ( i ) M

z
Known cross-
coupling
e ( i ) + eˆ ( i ) h ( ⋅) Noise estimator

C z
wd (i )
dˆu ( i ) Disturbance
Ld
dynamics
− D
z
wu ( i )
Lw
z

z
Reference y (i ) from
dynamics Reference plant/operator
du ( i ) M generator
Control law
u (i )

Embedded model

Control unit
104 E. Canuto et al.

The final expression relating ey to the model error e and to 3.2 Stability and performance versus structured and
the output disturbance dy in (4) is obtained as follows. unstructured uncertainty
Firstly, by replacing (15) and (18) in (47) one finds 3.2.1 Robust stability
ey = − ( I + S c ML w ) eˆ y − S c ML w e. (53) Input-output stability of (55) occurs if and only if:
Then, by replacing the prediction error eˆ y through (19) and 1 S is asymptotically stable, which in turn requires the
by defining the overall ‘control sensitivity’ S and the stability of Sm and Sc, and
complement V = 1 – S as follows: 2 the input signals e and dy are bounded and causally
S = Sm + Sw independent of ey .
V = 1 − S = Vm − S w (54) The latter condition only occurs when structured and
S w = S c ML wS m , unstructured uncertainties are zero: h = ∂P = 0 (the open
switches in Figure 2), i.e.,
the final expression is found to be
e( z ) = w y ( z )
e ( z ) = V ( z )e( z ) − S( z )d y ( z ). (55) (61)
d y ( z ) = M ( z ) ( D( z )w d ( z ) + wu ( z ) ) ,
The following alternative expression of V will be employed
hereafter: and imply that the tracking error is only forced by an
‘unpredictable’ noise as in the linear quadratic Gaussian
V = Vm − S c ML wS m = Vm − S c M ( H − DL d ) S m control (Kwakernaak and Sivan, 1972).
(56)
= Vm Vc + S c MDL d S m . On the contrary, making explicit structured and
unstructured uncertainties in (55) by means of (38) and (4),
The high-frequency asymptotes of the two terms in the last and dropping z, the implicit ‘error loop’ equation follows:
part of (56) descend from (8), (23) and (48), and hold
( )
ey = V∂P y − ey − SMh ( xc ) + Vwy
( f m∞ )m ∞
f c∞ ⎛ f ⎞
m∞ +1
(62)
lim f →∞< fmax Vm Vc = = ⎜ v∞ ⎟ −SM ( Dw d + wu ) .
f m∞ +1 ⎝ f ⎠ (57)
lim f →∞< fmax S c MDL d S m = ( f d ∞ / f ) Equation (62) is made explicit in ey through (34) and (36),
d∞
.
which yields the stability equation
Thus, if d∞ > m∞ + 1 and fv∞ ≥ fd∞, the former asymptote in
(57) dominates and the high frequency asymptote of V (1 + V∂P + SΔH ( x ) / φ ) ey = V ( ∂Py + wy )
simplifies to (63)
−SM ( h ( x ) + Dw d + wu ) .
m∞ +1
⎛ f ⎞
V∞ = lim f →∞< fmax V ( jf ) = ⎜ v∞ ⎟ , f ≥ f v∞ Equation (62) is graphically represented in Figure 2. There,
⎝ f ⎠ (58) the uncertainties in Figure 1 and the control unit are
f v∞ = κv∞ f m / γ1/ ( m∞ +1) , κv∞ = κ m∞ κc∞(
1/ m∞ +1)
. combined so as to build the ‘error loop’. The control unit,
detailed in Figure 7, has been transformed by the above
The overall functions S and V combine the predictor derivation into a combination of the ingredients of S and V,
transfer functions Sm and Vm with the causality correction namely, M (controllable dynamics), Sm (state predictor), Sc
Sw. The latter is imposed by wu not being rejected. Equation (control law and controllable dynamics) and Lw (noise
(55) is the basic ‘error loop’ equation, to be further estimator of wu).
elaborated by making explicit e and dy. Observe that:
1 the unstructured uncertainty ∂P is filtered by V, which
3.1.2 Example being a low-pass filter, implies that ∂P should become
C is a proportional plus derivative compensator, i.e., significant only at higher frequencies
C( z ) = c∞ ( z − 1) + c0 . (59) 2 the structured uncertainty ΔH is filtered by S, which
being a high-pass filter, implies that ΔH should become
Assuming that the poles of Sc are equal and collected in significant only at lower frequencies.
Λc = {1 – γc, 1 – γc}, the gain equations are c∞ = 2γc,
A sufficient stability condition in terms of the frequency
c0 = γc2 = κc 0 f c2 . The overall block diagram of the control
response may descend from the ‘small-gain theorem’ in
unit is in Figure 5. The errors Desoer and Vidyasagar (1975), and leads to
eˆθ (i ) = eθ (i ) + eˆθ (i )
(60) max f < fmax V ( f ) ∂P ( f ) + S( f )ΔH ( x , f ) / φ ( f )
eˆΔθ (i ) = eΔθ (i ) + eˆΔθ (i ) (64)
≤ η < 1,
are the ‘measured’ tracking errors already defined in (3).
Robust control stability using the error loop 105

where η-1 is a stability margin. Inequality (64), which is the S 0 ( jf ) = lim f →0 S( jf ) = lim f →0 S m (1 + S c ML w )
key result, is similar but not equal to inequalities in Doyle
= ( f / fs0 ) 0 , f ≤ fs0
d
et al. (1992), where stability and performance inequalities
are combined. (67)
f s 0 = f m 0 / (1 + lw0 / c0 )
1/ d 0

= κ m 0 f m / (1 + κ s γ nc )
1/ d0
3.2.2 Performance < fm0 ,

Inequality (64) allows to rewrite (63) as where κs = κw0 / κc0. Equation (67) affirms that the overall
sensitivity BW fs0 is narrower than the state predictor fm0,
ey ≤ (1 − η )
−1
( )
V ∂Py + wy − SM ( h ( x ) + Dw d + wu ) , (65) because Sc is not zero. Since, as stated in Section 2.4,
ΔH (x ) / φ is strictly proper and nc = degφ < d0, the
which becomes the performance inequality, |.| being a
suitable norm. It is out of the paper aim to discuss and apply parametric component SΔH / φ in (64) achieves its
(65). It suffices to point out that it can be split in two maximum close to fs0. Then, replacing the coefficients of
components: ΔH (x ) with their bound in (37), the following inequality
results
1 the ‘deterministic’ term V∂Py − SMh( x ) depending on
the reference trajectory max f < fmax S 0 ( f )ΔH ( x , f ) / φ ( f )
(68)
2 the ‘random’ term Vwy − SM (Dw d + wu ) depending ≤ (1 + εs ) ΔH max ( x, f s 0 ) / φ ( f s 0 )
on noise components. where εs << 1 accounts for the actual |S(f)| in the
Given η and the eigenvalue range derived from (64) in neighbourhood of fs0.
Section 3.3, the first term in (65) fixes the slew rate of the Maximisation of the second component in (64) employs
reference trajectory, whereas the random part fixes sensor the asymptote (58). Considering (39) and assuming f∂P ≥ fv∞,
and actuator noise. It has been already remarked that ey is a local maximum of |V∞(f)∂P(f)| occurs for fv∞ ≤ f ≤ f∂P,
only available through simulated runs, and is related to the since |V∞(f)| in (58) is monotonically decreasing. Having the
measured control error ey – defined as reference minus Example of Section 2.1.2 in mind, we assume
measure - through the model error e as in (3). Therefore, if max f < fmax V ( f )∂P ( f )
the tracking error ey is made bounded by (65) and (69)
≤ (1 + εv )( f v∞ / f ∂P ) ∞ ∂Pmax ( f ∂P ; p max ) ,
v
| ey | << | e | – which are mandatory requirements – the
control error ey becomes the opposite of the model error e. where εv < 1 plays the role of εs in (68).
Alternative and measurable conditions derive from the Finally, (64) splits into the asymptotic inequalities
former decomposition of ey in (3). In the case that | eˆy |→ 0,
( )
1/ v∞
which corresponds to approach the anti-causal law (52), the f v∞ ≤ f ∂P η / ∂Pmax ( f ∂P ; p max ) , η <1
(70)
control error becomes the opposite of the measured model
φ ( f s 0 ) ≥ ΔH max ( x, f s 0 ) / η, η < 1,
error e , a condition that can be easily verified in practice.
where εs in (68) and εv in (69) have been absorbed by the
3.3 Asymptotic stability inequalities stability margin 1/η.
Given η, (70) can be solved for fv∞ < fmax and fs0 < fmax.
Explicit expressions of (64) and (65) are obtained under the
The first inequality in (70) provides an upper bound to fv∞.
following assumptions.
The second is a polynomial inequality, that, since degφ >
1 State predictor and control law have uniform but degΔHmax(·), provides a lower bound as fs0 → ∞. Bounds in
distinct eigenvalues as already stated in (21) and (49). (70), if they are feasible, can be transformed into bounds on
Their ratio the state-predictor frequency fm, and on the control ratio γ in
(66), by replacing the expressions (67) and (58) of fs0 and
γ = γm / γc = f m / f c (66)
fv∞, respectively. At the end, (70) can be replaced by
is referred to as the control frequency ratio. f m,min (γ, η) ≤ f m ≤ f m ,max (γ, η) < f max . (71)
2 The components of the inequality (64) reach their
Inequality (71) proves the following.
maximum values within separate frequency domains:
the sensitivity frequency band, f ≤ fs0, and the 1 A robust design is driven by the predictor frequency fm,
complementary sensitivity band, f ≥ fv∞, to be defined which is bounded from above by the neglected
below. dynamics and bounded from below by the parametric
uncertainty.
Consider the former component in (64) due to parametric
uncertainty, and replace S(f) with the low-frequency 2 The admissible band Δfm = fm,max / fm,min ≥ 1 depends on
asymptote S0(f). The latter, derived from (22), (48) and (24), the control ratio γ > 0 and on the stability margin
holds
106 E. Canuto et al.

η–1 > 1. The asymptotic approximations that have been Figure 8 Admissible state predictor frequency fm for η = 0.5
adopted to obtain (71), tend to be accurate for γ ≤ 1. (see online version for colours)

Because of the assumptions made at the beginning of this 0.6 fm,max @ η=1/2
section, the closed-loop eigenvalues entailed by (71) must 10
fm,min @ η=1/2
be kept as first trial, to be refined by simulation and 0.5
10
in-field. The admissible band Δfm(γ) ≥ 1 provides
degrees-of-freedom to satisfy the performance inequality 0.4
10
(65) as in Canuto (2007, 2008).

fm [Hz]
0.3
10
3.4 Example: robust stability design
0.2
10
Using (9), (28) and (59), the low-frequency asymptote of S
and the high-frequency asymptote of V can be written as 0.1
10
S 0 ( z ) = ( z − 1) 4 β (1 + lu / c0 ) / ms
(72)
V∞ ( z ) = ( z − 1) −3 lu c∞ . 10
-1
10
0 1
10
φ [Hz/Hz]
They correspond to the asymptotes of the Bode plots in
Figure 6. Expressing fs0 and fv∞ in terms of fm and γ as in Figure 9 The admissible frequency band Δfm versus γ and η
(67) and (58), yields (see online version for colours)

(
f s 0 ( f m , γ ) = f m / 5 (1 + 2γ 2 ) )
1/ 4
η=0.2
(73)
f v∞ ( f m , γ ) = f m (20 / γ)1/3 . η=0.25
0 η=0.33
10
Δfm [Hz/Hz]

The stability inequalities in (70) become η=0.5


f v ( Δθ ) f g2
f s20 ( f m , γ ) − fs0 ≥
η η (74)
f v∞ ( f m , γ ) ≤ ft ,min ( 2ηζ t ,min ) .
1/3

They have been solved for fm and γ in the Table 1 using the
parameters of the same table. -2 0 2
10 10 10
Three alternative eigenvalues, denoted as cases 1, 2 and γ [Hz/Hz]
3, have been selected: they correspond to the square marks
on the line fm,max(γ) in Figure 8 and to γ = {1, 0.4, 0.2}. As The design has been restricted to the upper bound fm,max(γ)
shown by Figure 9, γ = 1 corresponds to the widest and to the region γ ≤ 1 in order to force the control error ey
admissible band Δfm, whereas γ = 0.2 is a lower threshold to stay within the target bounds in Figure 10. The frequency
since uncertainty cannot be more accommodated for γ < 0.2. bounds fm,max and fm,min are shown in Figure 8 versus the
In practice, given the state predictor frequency fm, the eigenvalue ratio γ for a stability margin η–1 = 2. The margin
control law frequencies fc and fv∞ in (73) become too large, is sufficient since the worst-case uncertainties enter the
and the first inequality in (70) (accommodating the bound expressions (74).
neglected dynamics) cannot be guaranteed.

Table 1 Parameters and design

Parameter Symbol Unit Value Comments


Gravity frequency fg Hz 0.11
Friction frequency fv Hz 0.25 Δθ ≥ 0.75 mrad

Transmission frequency ft,min Hz 22


Damping ζt,min 0.025
Backlash emin mrad ≤0.4
Predictor frequency: case 1 fm,max Hz 2.2 @ η = 0.5, γ = 1
idem, case 2 fm,max Hz 2.6 @ η = 0.5, γ = 0.4
idem, case 3 fm,max Hz 1.3 @ η = 0.5, γ = 0.2
Robust control stability using the error loop 107

Figure 9 shows the admissible frequency band Δfm. The slower in the set point achievement, as anticipated above.
horizontal dashed line corresponds to Δfm = 1 (the band has The backlash effect cannot be cancelled, being smaller than
no width). The largest achievable stability margin is close to the encoder quantisation. It couples with sensor quantisation
4. The widest admissible band occur for γ ≅ 1. and gives rise to a limit cycle around the set point not larger
than 1 bit.

4 Simulated and experimental results Figure 11 Set point achievement for two different designs
(see online version for colours)
The alternative designs in Table 1 have been tested through
simulation and a ball and beam device leaving the ball 2.002
Reference
uncontrolled. Their performance has been compared
Measure: γ=0.4

Angular position [rad]


through the control error ey defined in (3). The angular
position is measured by an incremental encoder mounted on 2.001 Measure: γ=0.2
the gear output shaft. Encoder quantisation is of the same
order of the backlash in Table 1. Up and down motion is
driven by a suitable reference generator that accounts for 2
speed and current limits, which latter impose motion
duration. Figure 10 shows the control error during up and
down motion and intermediate halt intervals. Bounds to 1.999
control error have been set larger during motion (dashed
lines in Figure 10). 4 5 6 7 8 9
Time [s]
Figure 10 Simulated control error for two different designs
(see online version for colours) The trapezoidal angular position of the gear pivoting the
-3 beam is measured by the gear encoder and is shown in
x 10
Figure 12. The difference between reference and measured
4 Tolerance
angle – the control error – can be perceived from the
Tolerance
3
enlargement in Figure 13, when the reference angle reaches
Control error: reference-measure γ=0.2
the maximum value. The measured position reaches the
Control error: reference-measure γ=0.4
2 maximum value after a damped oscillation ending in a limit
cycle (the square wave overlapping the constant reference)
Angle [rad]

1 imposed by backlash, friction and encoder quantisation.


Figure 12 shows the ball stroke (both simulated and
0 experimental), that, being uncontrolled, swings between the
beam extremes with some delay owing to friction.
-1
Figure 12 Experimental and simulated gear and ball motion
-2 (see online version for colours)
Gear Angular Positions
-3
2 4 6 8 10 12 14
Time [s] 0.8
0.6
Performance improves by increasing γ from 0.2 (case 3) to
Angular position [rad]

0.4 (case 2) and to 1 (case 1), as Figure 10 shows. The result 0.4
is reasonable since a lower γ in Figure 8 narrows the 0.2
predictor BW and consequently increases the norm of the
second term in the right-hand side of (65). This clearly 0
occurs during up and down motion, say from 9 s to 12 s in -0.2
Figure 10, when friction, assumed unknown, cannot be fully
-0.4 Experimental
cancelled by the predicted dˆu in (44). When the set-point is
-0.6 Simulated
reached, control error is dominated by backlash and no Ball stroke [m]: simulated
significant difference appears between the different cases. -0.8 Ball stroke [m] : measured
The case 1, not shown in Figure 10, behaves more or less as
the case 2, except that the control error becomes noisier, due 12 14 16 18 20 22
Time [s]
to a larger control BW. Therefore, case 2 looks a good
trade-off: the relevant transfer functions have been reported The transient oscillation in Figure 13 is the same as in
in Figure 6. Figure 11 and would disappear by adding an appropriate
Figure 11 shows the achievement of a set point for the known term h in the control law. To prove this, the
cases 2 and 3. One may observe that the case 3 is much
108 E. Canuto et al.

measured angular position of the simulated and Acknowledgements


experimental runs are compared in Figure 13. The
The paper is a revised and expanded version of a paper
experimental measurement is free of the transient oscillation
entitled ‘The error loop and robust closed-loop stability’,
because an appropriate h has been added to the control law
presented at IEEE ICMA 2012 Conference, Chengdu,
for accounting viscous and static friction. The result 5–8 August 2012.
highlights the different frequency bands where dˆu and
h (xˆ c ) counteract the plant uncertainty. Assuming that the
‘anti-causal limit’ holds, dˆu is estimated within the
References
sensitivity BW fm in (30). On the contrary, h (xˆ c ), Bickel, R. and Tomizuka, M. (1999) ‘Passivity-based versus
disturbance observer based robot control: equivalence and
depending on the controllable state xˆ c , is estimated within stability’, ASME J. Dynamic Systems, Measurement, and
the control law BW fc, which, in absence of command Control, Vol. 121, No. 1, pp.41–47.
limitations, may approach the Nyquist frequency Calafiore, G. and Dabbene, F. (2002) ‘A probabilistic framework
fmax = 0.5/T = 100Hz. The limit cycles in Figure 13 are for problems with real structured uncertainty in system and
different in their period because of a different static friction. control’, Automatica, Vol. 38, No. 8, pp.1265–1276.
Experimental measurements in Figure 13 certify that the Camacho, E.F. and Bordons, C. (2003) Model Predictive Control,
design model is a faithful description of the real plant. Springer Verlag, London.
Canudas de Wit, C., Olsson, H., Åström, K.J. and Lischinsky, P.
Figure 13 Gear set point achievement (enlargement of (1995) ‘A new model for control of systems with friction’,
Figure 12) (see online version for colours) IEEE Trans. on Automatic Control, Vol. 40, No. 3,
pp.419–425.
Gear Angular Positions
Canuto, E. (2007) ‘Embedded model control: outline of the
Experimental theory’, ISA Transactions, Vol. 46, No. 3, pp.363–377.
0.79 Simulated Canuto, E. (2008) ‘Drag-free and attitude control for the GOCE
satellite’, Automatica, Vol. 44, No. 7, pp.1766–1780.
Canuto, E. and Rolino, A. (2004) ‘Multi-input digital frequency
Angular position [rad]

0.785 stabilization of monolithic lasers’, Automatica, Vol. 40,


No. 12, pp.2139–2147.
Canuto, E., Acuna Bravo, W., Molano, A. and Perez, C. (2012)
‘Embedded Model Control calls for disturbance modeling and
0.78
rejection’, ISA Transactions, Vol. 51, No. 5, pp.584–595.
Canuto, E., Molano, A. and Massotti, L. (2010) ‘Drag-free control
of the GOCE satellite: noise and observer design’, IEEE
0.775 Trans. on Control Systems Technology, Vol. 18, No. 2,
pp.501–509.
Chapellat, H., Dahleh, M. and Bhattacharyya, S.P. (1990) ‘Robust
12 14 16 18 20 22 stability under structured and unstructured perturbations’,
Time [s] IEEE Trans. on Automatic Control, Vol. 35, No. 10,
pp.1100–1108.
Chen, S-H., Chou, J-H. and Zheng, L-A. (2005) ‘Stability
5 Conclusions robustness of linear output feedback systems with
time-varying structured and unstructured parameter
The error loop, i.e., the loop from model to tracking error, uncertainties as well as delayed perturbations’, J. Franklin
points out which is the key design tool for accommodating Institute, Vol. 342, No. 2, pp.213–234.
uncertainties. The tool is the noise estimator, which is Choi, Y., Yang, K., Chung, W.K., Kim, H.R. and Suh, I.H. (2003)
responsible for the noise estimates that update the ‘On the robustness and performance of disturbance observers
disturbance state in the embedded model. Actually, for second-order systems’, IEEE Trans. on Automatic
causality constraint adds a further degree-of-freedom to the Control, Vol. 48, No. 2, pp.315–320.
design, allowing the control law to play a role in making the Desoer, C. and Vidyasagar, M. (1975) Feedback Systems:
design feasible, especially when contrasting stability and Input-Output Properties, Academic Press, New York.
performance inequalities need to be satisfied. Employing Donati, F. and Vallauri, M. (1984) ‘Guaranteed control of
asymptotic expansions of the closed-loop transfer functions, ‘almost-linear’ plants’, IEEE Trans. on Automatic Control,
Vol. 29, No. 1, pp.34–41.
simple and explicit design formulae relating closed-loop
eigenvalues to model parameters and requirements have Doyle, J.C., Francis, B.A. and Tannenbaum, A.R. (1992) Feedback
Control Theory, Macmillan Pu. Co., New York.
been derived. Refinement may be pursued through
simulation and in field.
Robust control stability using the error loop 109

Foo, Y.K. and Soh, Y.C. (1993) ‘Robust stability bound for Mita, T., Hirata, M., Murata, K. and Zhang, H. (1998) ‘H∞ control
systems with structured and unstructured perturbations’, IEEE versus disturbance-observer-based control’, IEEE Trans. on
Trans. on Automatic Control, Vol. 38, No. 7, pp.1150–1154. Industrial Electronics, Vol. 45, No. 3, pp.488–495.
Hostetter, G. and Meditch, J. (1973) ‘On the generalization of Patre, P.M., MacKunis, W., Makkar, C. and Dixon, W.E. (2008)
observers to systems with unmeasurable, unknown inputs’, ‘Asymptotic tracking for systems with structured and
Automatica, Vol. 9, No. 6, pp.721–724. unstructured uncertainties’, IEEE Trans. on Control Systems
Johnson, C. (1971) ‘Accommodation of external disturbances in Technology, Vol. 16, No. 2, pp.373–379.
linear regulator and servomechanism problems’, IEEE Trans. Ross Barmish, B. (1994) New Tools for Robustness of Linear
on Automatic Control, Vol. 16, No. 6, pp.635–644. Systems, Macmillan Pu. Co., New York.
Katsura, S., Matsumoto, Y. and Ohnishi, K. (2007) ‘Modelling of Schrijver, E. and van Dijk, J. (2002) ‘Disturbance observers for
force sensing and validation of disturbance observer for force rigid mechanical systems: equivalence, stability, and design’,
control’, IEEE Trans. on Industrial Electronics, Vol. 54, ASME J. Dynamic Systems, Measurement, and Control,
No. 1, pp.530–538. Vol. 124, No. 4, pp.539–548.
Keshmiri, M., Jahromi, A.F., Mohebbi, A., Amoozgar, M.H. and Solihin, M.I., Akmeliawati, R. and Legowo, A. (2011) ‘Robust
Xie, W-F. (2012) ‘Modelling and control of a ball and beam feedback control design using PSO-based optimization: a case
system using model based and non-model based control study in gantry crane control’, Int. J. of Mechatronics and
approaches’, Int. J. on Smart Sensing and Intelligent Systems, Automation, Vol. 1, No. 2, pp.121–131.
Vol. 5, No. 1, pp.14–35. Wu, B-F. and Jonckheere, E.A. (1992) ‘A simplified approach to
Kwakernaak, H. and Sivan, R. (1972) Linear Optimal Control Bode’s theorem for continuous-time and discrete-time
Systems, J. Wiley & Sons, New York. systems’, IEEE Trans. on Automatic Control, Vol. 37, No. 11,
Maciejowski, J. (1989) Multivariable Feedback Design, Addison pp.1797–1802.
Wesley, Wokingham, Berskshire, UK.

You might also like