PhDThesis ChihFengLee Web
PhDThesis ChihFengLee Web
compensation of an automotive
electromechanical brake
July 2013
All rights reserved. No part of this thesis may be reproduced in any form or by any means,
without permission in writing from the author.
Typeset in LATEX 2ε
This thesis is printed on archival quality paper
Melbourne, 2013
Abstract
iii
iv
imental results show that the rise time of the NTOC is typically within 5%
of the time-optimal control without control chatter. The NTOC also demon-
strates significantly higher bandwidths than the industry-tuned PI controller,
which makes it viable for advanced algorithm deployment.
Furthermore, a holistic control design approach that considers tracking
error, control effort and system constraints is examined under the model pre-
diction control (MPC) framework. This design approach exhibits greater flex-
ibility and slight control performance improvements compared to the NTOC.
However, the intensive computational and large memory demands are pro-
hibitive under current hardware limits.
By utilising the high bandwidth clamp force tracking control of the NTOC,
a novel method for attenuating cold judder directly from the source of vibra-
tion is proposed. Two compensator designs are developed, whereby the EMB is
utilised to generate a compensating clamp force to counteract the judder caus-
ing brake torque variation (BTV). The first approach is the linear parameter-
varying (LPV) BTV compensator, where it is scheduled using the measured
wheel speed and estimated acceleration. Experimental investigations show fa-
vorable results, including a first-order implementation reducing BTV by up to
55%.
To mitigate the limitations arising from relatively low sampling rates, an
alternative approach examines adaptive feedforward compensation, where the
compensator is scheduled using the wheel position, speed and acceleration. Ex-
perimental investigations demonstrate improved BTV attenuation compared
to the LPV compensator, albeit with the additional requirement of wheel po-
sition measurements.
Declaration
This is to certify that:
1. the thesis comprises only my original work towards the PhD except where
indicated,
2. due acknowledgement has been made in the text to all other material
used,
3. the thesis is less than 100,000 words in length, exclusive of tables, maps,
bibliographies and appendices.
v
Acknowledgements
First and foremost, I would like to express my deep gratitude to my supervisor,
Associate Professor Chris Manzie for accepting me into the group, patient
guidance, enthusiastic encouragement, continuous support and useful critiques
throughout the project.
I would like to thank Professor Malcolm Good, who headed the Melbourne
University node of the Research Centre for Advanced By-Wire Technologies
(RABiT) for providing a pleasant working environment. I wish to acknowledge
the help provided by the engineers at Pacifica Group Technologies, Sam Oliver,
Tony Rocco and Peter Harding for giving technical support and troubleshoot-
ing tips in regards to the control implementation on the electromechanical
brake (EMB), and to the technician of the Advanced Centre for Automotive
Research and Testing (ACART), Don Halpin for his excellent machining skill.
I would also like to thank Chris Line for sharing his parameter identifica-
tion script. The EMB parameter identification script used in Section 3.2 was
based on the work of Chris Line, contributed by Caroline Jekelius (a master ex-
change student from TU Munich working under my supervision in Melbourne
University), and ported to Vector CANape scripting language by myself.
I am indebted to my undergraduate supervisor, Dr Janusz Krodkiewski
for introducing me to the joy of research during my undergraduate summer
research project and final year project. I am also grateful to Professor Dragan
Nešić, Rahul Sharmar, Alan Chang and Sei Zhen Khong for various discussions
that have broaden my understanding in control theory and mathematics in
general. Many thanks to Professor Zhihong Man, Associate Professor Edwin
Tan and Mark Ng for their cordial reception of my visit to Monash University
in Malaysia.
Special thanks should be given to my hyung in the Dynamics and Control
Group, Brett Bishop and Tae Soo (Terence) Kim, and to my contemporaries
in the group, Alireza Mohammadi, Jalil Sharafi, Kuan Waey Lee, Michael
Stephens and Ronny Kutadinata for their assistance and encouragement at
vii
viii
different phases of the project. In particular, I would like to thank Terence for
enjoyable and inspirational soju sessions. I wish to acknowledge the friendship
and encouragement of Chuck Yong Kong, Darwin Lau, Sei Zhen Khong, Thi-
laksiri Bandara, Tien Mun Foong, Xin Fu Tan, Yee Khoon Tee and Yee Wei
Law. I would also like to express my very great appreciation to Ming Hui Siew
for her love, encouragement and support.
Last, but certainly not least, I wish to thank my parents and sister for their
guidance, encouragement and support. My mother and sister have been cook-
ing delicious meals for me and my mechanical engineer father has demonstrated
to me that choosing engineering as a career can both be fun and rewarding.
A Graduate Certificate in Commercialisation has been completed during
the course of this work, and is subsidised under the Federal Government’s
Commercialisation Training Scheme (CTS). The PhD research has been fi-
nancially supported by the Australian Postgraduate Award and Special Stu-
dentships from the Mechanical Engineering Department. Travel is funded by
the Melbourne School of Engineering Research Training Conference Assistance
Scheme.
Contents
1 Introduction 1
1.1 Brake-by-wire systems . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Brake judder . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
ix
x Contents
3.2.3 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . 60
3.3 Development of a control-oriented EMB model . . . . . . . . . . 63
3.4 Modelling of the characteristics of brake torque variation . . . . 68
3.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
8 Conclusions 201
8.1 Contributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202
8.2 Directions for further work . . . . . . . . . . . . . . . . . . . . . 203
Bibliography 211
List of Figures
1.1 Growing pace of the electronics content of vehicles, with ongoing
adoption of by-wire technologies. . . . . . . . . . . . . . . . . . . . 2
1.2 Brake-by-wire system. . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 Cutaway view of the PGT electromechanical brake. . . . . . . . . . 5
1.4 Disc runout and disc thickness variation. . . . . . . . . . . . . . . . 7
1.5 Brake judder transfer path. . . . . . . . . . . . . . . . . . . . . . . 8
1.6 Strategies for brake judder attenuation. . . . . . . . . . . . . . . . . 9
xii
List of Figures xiii
5.27 Comparison of rise time and energy usage of the time-optimal con-
trol, NTOC and MPC. . . . . . . . . . . . . . . . . . . . . . . . . . 149
5.28 Responses of 20 kN step using the time-optimal control, NTOC and
MPC. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
5.29 EMB closed-loop system and clamp force controllers developed in
this chapter. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
xvii
Notations
Acronyms
ABS Anti-lock braking system
BBW Brake-by-wire
BPV Brake pressure variation
BTV Brake torque variation
CAN Controller area network
EHB Electrohydraulic brake
EKF Extended Kalman filter
EMB Electromechanical brake
EMF Electromotive force
ESC Electronic Stability Control
EWB Electronic wedge brake
FFT Fast Fourier transform
LMI Linear matrix inequality
LPV Linear parameter-varying
LTV Linear time-varying
MIMO Multiple-input multiple-output
MPC Model predictive control
NTOC Near-time-optimal control
PMP Pontryagin’s minimum principle
PMSM Permanent Magnet Synchronous Motor
RMSE Root Mean Square Error
SISO Single-input single-output
xix
xx Notations
Functions
Diagonal matrices:
0
A1
diag(A1 , . . . , An ) =
...
(1)
0 An
Sign function:
1,
if x > 0,
sgn(x) = 0, if x = 0, (2)
−1, if x < 0,
Symbols
θd Disc rotor angular position
θm EMB motor angular position
θm,r EMB motor angular position reference
ωm EMB motor angular velocity
ωm,r EMB motor angular velocity reference
ωd Disc rotor angular speed
ωs Characteristic velocity of the Stribeck friction
µ Coefficient of friction between brake pads and rotor
µs Coefficient of load-dependent static friction
H Hamiltonian
L Lagrangian
*Symbols that are not listed here are defined close to its use in the body of the thesis.
Publications
The material presented in this thesis is based on a number of refereed papers,
which have appeared during the course of my PhD research. In particular,
Chapter 4 is based on:
xxiii
xxiv Publications
Introduction
1
2 Chapter 1. Introduction
Electrical/Electronics Content of Vehicles (%)
• Keyless vehicle
30 • Occupant sensing
• Rollover sensing
• Remote keyless entry
• Integrated powertrain control
25 • Electronic instrument cluster • Multiplexing
• Traction control • Diagnostics • Throttle-by-wire
• Adaptive suspension • Navigation systems • Automated parking
20 • Crash sensor diagnostics
• Theft deterrent systems • Adaptive cruise control
• Stability control
• Cruise control
15 • Side impact sensor
• Electronic fuel injection
• Tyre pressure sensor
• Electronic transmission control
• Single point crash sensor
• Lamp
10 • Horn • Breakerless ignition • Vehicle immobiliser
• Alternator • Electrical power steering
• Radio
• Generator • Trip computer
5 • Starter • Service indicator
• Antilock braking
• Cellular phone
0
1950 1960 1970 1980 1990 2000 2010
Year
Figure 1.1: Growing pace of the electronics content of vehicles, with ongoing
adoption of by-wire technologies (adapted from Winters et al. (2004)).
the trend observed in the last few decades, whereby a steady increase of electri-
cal, electronics and mechatronics content in vehicles, augmenting or replacing
the mechanical components that had been traditionally dominating the share
of vehicle parts (Figure 1.1). The major adoption of advanced vehicular elec-
tronics began in the 1970’s in order to comply with the Clean Air Act in the
USA (passed in 1963 and amended in 1970). This motivated the use oxygen of
sensors and powertrain control technologies to reduce emission pollution and to
mitigate the smog problem (Klier and Rubenstein, 2011). More recently with
the introduction of more stringent emission standards, increasing demands of
fuel efficiency and high engine performance, the throttle-by-wire system has
became a standard feature in many vehicles (Bretz, 2001).
To fully realise the benefits of drive-by-wire requires the integration of three
primary vehicle control systems, namely throttle, steering, and braking. Steer-
ing system technology has evolved from hydraulic to electric power-assisted
steering, whereby assisting torque is supplied by an electric motor to reduce
the effort required to turn the wheels. Modern advancements including the
speed sensitive steering and automated parallel parking are also available in
many production vehicles. The full steer-by-wire provides the greatest design
1.1. Brake-by-wire systems 3
PBS CCU
3 3
6 5 4
ments for motor and power sizings. However, the amount of actuator force
that can be exerted by the motor without self-energising may not be sufficient
to use as a park brake. On the other hand, the conventional EMB is the more
popular design, whereby many automotive Original Equipment Manufactur-
ers (OEMs), including Bosch (Schumann, 2001), Continental (Halasy-Wimmer
and Völkel, 2005), Delphi (Siler and Drennen, 2002) and Hitachi (Watanabe
et al., 2007) have their respective models. This thesis employed the EMB de-
signed and manufactured by Pacifica Group Technologies (PGT) (Wang et al.,
2005). It comprised an electric motor, a gear reduction stage, a ball screw, a
floating calliper and control electronics (Figure 1.3). Its main features include
continuous brake force actuation and near elimination of the residual drag1 .
Compared to the hydraulic braking systems, the absence of brake booster,
vacuum pump and hydraulic lines in the EMB-based BBW systems leads to
the reduction in component number, installation space, assembly time and
maintenance costs. Environmental benefits are achieved by avoiding brake
fluid usage, which is corrosive and toxic.
Furthermore, the pedal no longer vibrates during ABS operation, thus
lessening the risk of an inexperienced driver erroneously reducing the brake
pressure and prematurely terminating the ABS control mode. Other safety
features include redundant wiring, extra batteries to offset current peaks and
lessen the risk of power failure under a defective generator2 , and fault-tolerant
1
Brake pads contacting with the disc during off-brake running.
2
In line with the regulatory requirements to split the hydraulic brake systems into two
1.2. Brake judder 5
1 Calliper, 1 2 3 4 5 6
2 Ball screw shaft,
3 Motor rotor,
4 Motor stator,
5 Planetary gears,
6 Control board.
mid to high speed (30 to 140 km/h) (Jacobsson, 2001; Lee and Brooks, 2003).
Brake judder is defined as a braking-induced forced vibration occurring at the
steering wheel, brake pedal or vehicle body, where the frequency of vibration
is proportional to the wheel speed3 (Jacobsson, 2001). It has resulted in major
recalls, warranty claims (Nissan North America, 2005) and detrimental effects
to brand image. Furthermore, the associated vibrations also induce additional
wear and prompt components fracture, hence resulting in shorter service life
and higher operating costs.
Brake judder is caused by the irregularities in the brake disc, and can be
categorised into cold and hot (or thermal) judder (Abdelhamid, 1997). Cold
judder is induced by geometrical irregularities due to mounting, machining,
uneven wear, uneven corrosion and uneven friction film generation (Jacobs-
son, 2003a). The irregularities associated with cold judder occur fewer than
five times per disc revolution, and correspond to vibration frequencies which
are lower than five times of the wheel rotational speed (Jacobsson, 2001). Geo-
metrically, these irregularities resemble disc runout or disc thickness variation
(DTV) (Figure 1.4). Disc runout may induce uneven wear and further gen-
erates DTV (Haigh et al., 1993a), which contributes to the brake pressure
variation (BPV) and brake torque variation (BTV). For most manufacturers,
the manufacturing tolerances for disc runout and DTV are less than 80 µm
and 10 µm respectively (de Varies and Wagner, 1992). However, a slight in-
crease in DTV to 15 µm can cause brake judder in vehicles that are sensitive to
judder (Jacobsson, 2003a), therefore raising the need for better manufacturing
tolerances.
Hot judder is induced by thermal deformation (e.g. coning and waving),
uneven thermal expansion, or phase transformation of the disc material (Jacob-
sson, 2003a). Unlike cold judder, hot judder does not require large amplitude
permanent disc irregularities for its initialisation. When exposed to braking
pressure, permanent irregularities (both with low and high amplitudes) cause
a localised rise in contact pressure field and temperature, leading to hot spots
generation (Kubota et al., 1998; Little et al., 1998; Jacobsson, 2001; Sury-
atama et al., 2001). The hot spots resulted in instantaneous DTV and spatial
friction variation, causing brake pressure and brake torque pulsations during
3
It is usual to relate judder frequencies to wheel speed, e.g. frequency at twice the
number of wheel speed is called second order judder.
1.2. Brake judder 7
1 Brake pad, 1 2
2 Brake disc,
3 Wear due to pad rubbing. Disc Disc
2 Rotation Rotation
3 Rotation
Axis
1
3
Excitation
Brake disc irregularities
BPV BTV
Transfer
Hydraulic system Wheel suspension & Amplification
Amplification
Vibration Car body Steering system & Perception
• Improved
manufacturing
tolerance
• Resurface disc
• Improved heat
transfer
explored. While the brake pedal feel emulator equipped in BBW system is
capable of eliminating pedal pulsation due to BPV, the resultant BTV that
causes judder is not dealt with. It is still uncertain whether the additional
bandwidth offered by EMBs can be used to actively attenuate the BTV. This
would reduce the need for a better manufacturing tolerance and extend service
life of components.
2
11
12 Chapter 2. Literature review and research aims
Fcl
Fcl,r Force ωm,r Speed iq,r Current EMB
+ Controller + Controller + Controller Actuator
− − −
Fcl ωm iq
Figure 2.1: EMB clamp force control with cascaded control structure (Schwarz
et al., 1998).
with the speed controller generating the current reference. The innermost loop
contains the current controller, where the motor voltage is calculated and ap-
plied to the motor via a Pulse Width Modulation (PWM) driven inverter. The
purpose of the current controller is to deal with the dynamics of the power sup-
ply and the armature, and to cancel the effect of the induced armature voltage
or Back Electromotive Force (back EMF). Since the motor torque is propor-
tional to the current, the current loop can also be called the torque loop. Note
that this multi-loop control structure requires the bandwidth of the control
increases towards the inner loops with the current loop being the fastest and
the force loop the slowest. This is to ensure reference generated by the outer
loop can be handled well by the inner loop (Leonhard, 2001).
To address the motor speed limit, Maron et al. (1997) design the force con-
troller using a proportional controller with output saturation, and the velocity
controller is implemented using a PI controller. While the presence of current
controller is inevitable in implementation, it is omitted in the model presented
this paper, and the measured motor current is assumed to be identical to the
reference. This assumption is based on the fact that the current loop has a
relatively fast dynamics compared to the dynamics of the mechanical compo-
nents.
Similar work was done by Schwarz et al. (1998), albeit with more details on
the current controller design. Likewise, the force controller is a proportional
controller implemented on a microcontroller. However, the speed and current
controllers are both PI controllers with series-connected first-order lag elements
implemented using analog circuits. The task of the first-order lag elements
are to attenuate measurement noise. Additionally, the motor current limit is
addressed in this work by the inclusion of an output saturation component in
the speed controller.
Note that compared to the standard motion control problem, the EMB
clamp force control is differentiated by a large operating range in which non-
14 Chapter 2. Literature review and research aims
linear friction and nonlinear stiffness become significantly high at large clamp
force (of up to 30 kN). In order to improve the dynamic response, Line et al.
(2004) suggest augmenting a static friction compensation to the current refer-
ence. The compensating current reference introduces extra motor current in
order to break away the stiction. This improvement reduces the tracking error
when the motor velocity switches sign and crosses zero.
In a follow up paper, Line et al. (2006) show that the dynamic response can
be further improved by applying feedback linearisation techniques. Given the
block diagram of the implementation shown in Figure 2.2, compensating cur-
rent reference is augmented to cancel the nonlinear response due to Coulomb
friction as well as nonlinear stiffness. This technique transforms the nonlinear
EMB actuator system into an equivalent linear system (as seen by the cas-
caded PID controller), hence allows for to a more intuitive gain tuning. The
use of gain scheduling for the force and velocity controllers is also suggested
to enhance tracking performance over the entire working envelop of the EMB.
Nevertheless, scheduling the entire set of gains in real-time increases memory
and processing requirements. By taking a compromising step, Jo et al. (2010)
tackles this impediment by scheduling only the proportional gain in the force
controller, and the derivative and integral gains remain fixed.
Despite their simplicity, the cascaded PID control structure has several lim-
itations associated with high performance clamp force tracking of EMB. The
key limitations of cascaded PID controllers associated with this research are:
control in the presence of nonlinear friction and nonlinear stiffness, maintaining
performance against actuator saturations, and tuning for optimal performance.
Although techniques to include auxiliary mechanisms for controller integration
have been developed, they appear to be ad-hoc additions to the standard cas-
caded PID controller and the optimal performance is not guaranteed. Utilising
a suitable model of the plant that is to be controlled can assist with the design
of a high performance controller, and this approach will be reviewed in the
following subsection.
iq,f Feedback
Linearisation
+ Fcl
Fcl,r Force ωm,r Speed iq,r Current EMB
+ Controller + Controller + Controller Actuator
− − −
Fcl ωm iq
Figure 2.2: EMB clamp force control with cascaded control structure and
feedback linearisation (Line et al., 2006).
designing a satisfactory control structure and tuning the gains. The control
design paradigm that utilises a model is called model-based control design.
Holweg et al. (2000) make use of the plant model to develop an inverse-
model controller, where nonlinear terms can be dealt with over the whole
operating region. However, since the controller is the inverse of the model,
the controller only performs well if the model is accurate. Generally, there is
no model that can exactly represent the plant and modelling discrepancy is
inevitable. Furthermore, the plant dynamics are likely to change with time due
to factors such as wear of mechanical components and thermal effects. When
inverse models are used in controller synthesis, even the slightest amount of
modelling inaccuracy can yield significant steady state error. In other words,
an inverse-model controller is not robust against modelling error.
To increase the robustness of the controller against modelling error, Krish-
namurthy et al. (2005, 2009) propose the application of backstepping design
method in EMB control. While the controller is analytically shown to be robust
to bounded parametric uncertainties, the design method lacks a systematic
routine for picking the controller gains to achieve good tracking performance.
Another approach to handle modelling error is offered by Line et al. (2007),
in which the parametric uncertainty is dealt with using linear robust control.
Figure 2.3 displays the block diagram of the proposed control scheme. Com-
pared to the cascaded PID control structure, it can be noted that the linear
robust controller has now taken over the role of force and speed control loops.
The controller synthesis is first started with the implementation of feedback
linearisation in order to transforms the nonlinear plant model into an equiv-
alent linear model. This is followed by gain selection for the linear robust
controller. Note that a systematic gain tuning procedure is available based on
the H∞ control theory (Zhou and Doyle, 1998). Since the controller is synthe-
16 Chapter 2. Literature review and research aims
iq,f Feedback
Linearisation
+ Fcl
Fcl,r Linear Robust iq,r
Controller Current EMB
+ Controller Actuator
(or MPC)
−
iq
Fcl ωm
Figure 2.3: EMB clamp force control with linear robust controller or model
predictive control with feedback linearisation (Line et al., 2006, 2007).
sised upon the feedback linearised model, the set of gains is only optimal with
respect to the linear model. In other words, the optimal performance cannot
be guaranteed for the original plant model. Additionally, the input and state
saturations are not explicitly considered during the synthesis. Based on simu-
lation and experimental results, the linear robust control is found to be overly
conservative, where the tracking performance is sluggish (Line, 2007).
Notice that backstepping control and robust linear control contain fixed sets
of controller parameters which are synthesised to be robustly stabilising for a
bounded set of parametric uncertainty. Kwak (2005) takes the idea of robust
control further, where the robust controller is augmented with an auxiliary
mechanism to adaptively tune the controller parameters online. While this
presents an interesting concept, the controller structure is complex and requires
differentiation of measured values — all of these may lead to difficulties in
implementation1 .
To address the sluggish performance of the linear robust control, robustness
is compromised by replacing it with a linear unconstrained model predictive
control (MPC) (Line et al., 2008). MPC employs a model of the plant to
predict how the system will respond to a set of control inputs, and then op-
timises these inputs so that the predicted output closely coincides with the
reference. A notable advantage of the linear unconstrained MPC is that a
systematic synthesis procedure can be applied offine to optimally choose the
controller parameters (Maciejowski, 2002). Furthermore, this control struc-
ture resembles that of a state feedback controller with simple implementation.
However, similar to the linear robust controller presented previously, the in-
put and state constraints are not explicitly considered during the controller
synthesis. To handle these constraints, post-saturation with varying limits is
1
Experimental results are not included in this work.
2.1. Electromechanical brake control 17
applied on the controller output during implementation. Apart from the ad-
hoc constraints management, optimality of the controller with respect to the
original nonlinear plant model is also not guaranteed. Note that additional
motor current is demanded by the feedback linearisation block. Hence, the
current control available to the MPC is reduced in the face of current limit.
Since the controller is optimally tuned for the feedback linearised model, it
may not perform optimally on the original plant. Nevertheless, experimental
results demonstrate that this controller is simple to implement with improved
performance compared to the cascaded PID control.
where t0 , tf are the initial and final time respectively; and g, h are scalar func-
tions. As usual, x, u are the plant state and input respectively. The problem
of finding a control policy u(·) that optimises the performance measure (2.1)
is called the optimal control problem.
In many applications, such as robot arm control, automatic assembly sta-
tions and disc drive heads, controllers are designed to make optimal use of the
maximum torque available for rapid maneuvers (Newman and Souccar, 1991;
Zhang, 2000). Rapid clamp force reference tracking in EMB is also desired
for a responsive braking response and shorter braking distance. The require-
ment of time minimisation are considered in the time-optimal control design,
whereby the aim is to find a control policy that transfers the system state from
an arbitrary initial condition, x(t0 ) = x0 to a specified target set S in minimum
time. If the target set S is given by a reference, xr (·), then the performance
measure to be minimised is given by
Z tf
Jto := tf − t0 = 1 dt,
t0
where tf is the first instant of time when the trajectory of x coincides with
xr (·).
To facilitate the following discussions in this section, a general nonlinear
plant model is considered:
ẋ(t) = f x(t), u(t), t ,
2.2. Time-optimal control 19
where x(t) ∈ X ⊆ Rnx and u(t) ∈ U ⊆ Rnu . Furthermore, the assumption that
X and U are compact allows state and input constraints to be considered. This
is commonly applicable in practical situations as state and input constraints
are usually found in actual plants. The time-optimal control for the general
nonlinear plant model with constraints can be formulated as:
Z tf
min 1 dt, (2.2a)
u(t) t0
subject to ẋ(t) = f x(t), u(t), t , (2.2b)
x(t) ∈ X , (2.2c)
u(t) ∈ U, (2.2d)
x(t0 ) = x0 , (2.2e)
x(tf ) = xr (tf ). (2.2f)
for all t ∈ [t0 , tf ] and p ∈ Rnx denotes the costate vector. Note that optimal
terms will be superscribed by ∗ in the following. According to Pontryagin’s
minimum principle (PMP) presented in Kirk (2004), necessary conditions for
the time-optimal control, u∗ (t) are
H x∗ (t), u∗ (t), p∗ (t), t ≤ H x∗ (t), u(t), p∗ (t), t (2.4)
and
− p∗ (tf )T δxf + H x∗ (tf ), u∗ (tf ), p∗ (tf ), tf δtf = 0. (2.5)
h i
dx (t )
It should be pointed out that δxf = 0 for fixed xr , while δxf = r f
dt
δtf for
time-varying state reference xr (t). Additionally, the optimal state and costate
are governed by
∂H ∗
ẋ∗ (t) = x (t), u∗ (t), p∗ (t), t , (2.6a)
∂p
∂H ∗
ṗ∗ (t) = − x (t), u∗ (t), p∗ (t), t , (2.6b)
∂x
and the boundary conditions are given by
Furthermore, if the Hamiltonian does not explicitly depend on time, then the
following additional necessary condition can be shown:
dH ∗
x (t), u∗ (t), p∗ (t) = 0, t ∈ [t0 , tf ]. (2.8)
dt
Note that PMP only provides necessary condition for optimality. For cases
where input is unbounded, or the bounds on input are conceived as arbitrar-
ily large, the following additional necessary and sufficient conditions for local
maximum principle (Pontryagin et al., 1962), and this resulted in the equivalent Pontryagin’s
minimum principle.
2.2. Time-optimal control 21
∂H
x∗ (t), u∗ (t), p∗ (t), t = 0,
∂u
∂ H
2
x∗ (t), u∗ (t), p∗ (t), t > 0.
∂u2
This result can be extended to the global optimality if the Hamiltonian is
convex and can be expressed in the form
H x(t), u(t), p(t), t
T 1
= c x(t), p(t), t + m x(t), p(t), t u(t) + uT (t)R(t)u(t),
2
where m(·) does not have any terms containing u(t).
where the input constraint and state constraint are given respectively by
fsu u(t), t ≤ 0, (2.9a)
fsx x(t), t ≥ 0, (2.9b)
and µu (·), µx (·) are the multipliers. Using the newly defined Lagrangian, the
costate dynamics (2.6b) are modified as
∂L ∗
ṗ∗ (t) = − x (t), u∗ (t), p∗ (t), t .
∂x
22 Chapter 2. Literature review and research aims
∂L ∗
x (t), u∗ (t), p∗ (t), t = 0,
∂u
µu (t)fsu u∗ (t), t = 0, µu (t) ≥ 0
µx (t)fsx x∗ (t), t = 0, µx (t) ≥ 0
dH ∗ dL ∗
x (t), u∗ (t), p∗ (t), t = x (t), u∗ (t), p∗ (t), t
dt dt (2.10)
∂L ∗
= x (t), u∗ (t), p∗ (t), t .
∂t
Clearly, this approach is not robust against disturbance and is likely to yield
large discrepancy from the intended optimal trajectory.
Another approach is to determine the number of control switching (Pon-
tryagin et al., 1962; Sussmann, 1979) and to compute the associated switching
function, which the latter provides the instant of control switching as a func-
tion of the state. A switching curve that is computed prior to online imple-
mentation facilitates the construction of a feedback control structure. As an
example, the switching function for the time-optimal control that regulates a
double integrator system:
ẋ1 = x2
(2.11)
ẋ2 = u, |u| ≤ 1,
to the origin can be analytically calculated as
q
fto (x1 ) = − sgn(x1 ) 2|x1 |, (2.12)
where the optimal state trajectories using control law (2.13) for several initial
values are also shown in Figure 2.4.
Implementation of time-optimal control is challenged by several issues.
First, the time-optimal control changes sign across the switching curve, resem-
bling a control with infinite gain. The sharp transition of control magnitude
may excite structural modes during maneuvers (Albassam, 2002). Further-
more, the switching function is sensitive to modelling errors, whereby impos-
ing an arbitrarily small perturbation to the model can lead to control chatter
(Ryan, 1980). In addition to modelling uncertainties, feedback delays and mea-
surement noises also induce chatter. The resulting chatter leads to vibration,
excessive heat lost in power electronics and wear in mechanical components
— all clearly undesirable in a braking application. Figure 2.5 demonstrates
the occurrence of control chatter in the double integrator system (2.11) when
the state is in vicinity of the origin. Third, apart from some simple low-order
systems, the control switching function is in general difficult to obtain and
usually does not have a analytic expression.
24 Chapter 2. Literature review and research aims
10
B u*=−1
u*=+1
5
u* = −1
dx/dt
−5
A
u* = +1
−10
−25 −20 −15 −10 −5 0 5 10 15 20 25
x
3
Control Switching Control Chatter
2
1
u
−1
−2
0 1 2 3 4 5 6 7 8 9 10
0
x, dx/dt
−5
x
dx/dt
−10
0 1 2 3 4 5 6 7 8 9 10
Time (s)
Figure 2.5: Control and state trajectories of the double integrator system
(2.11) with initial condition at (−10, 0.5). Control chatters when the state is
in vicinity of the origin.
2.2. Time-optimal control 25
10
TOC
5
u = −1
dx/dt
−5
u = +1
−10
−25 −20 −15 −10 −5 0 5 10 15 20 25
x
where the parameters α, ks , ssat and x2 sat can be tuned according to the guide-
lines provided in (Newman, 1990). This algorithm comprises two operating
range, a reaching phase where maximum allowable control is used, and a
boundary layer close to the manifold when the control is not saturated (Fig-
ure 2.6). Furthermore, it reduces the switching structure with two cases found
in the original time-optimal control law (2.15) to a single case only, there-
fore is simpler to implement. Another novelty is presented when the state
approaches the origin, the controller becomes a proportional-derivative (PD)
controller with a velocity-squared dissipative term:
αks x1 αx2 αks x2 |x2 |
untoc,unsat. = − − − (2.17)
ssat x2 sat ssat
The derivative term 1/x2 sat can be adjusted to reduce the magnitude of over-
shoot. This approach results in a very high performance control suitable for
high speed tracking with simple tuning and implementation, while free of chat-
ter.
Further extension of the NTOC is presented in (Newman and Souccar,
1991), where a nonlinear second-order system is considered. Additionally, the
expression of the manifold (2.14) is generalised. For example, (2.14) can be
modified such that its curvature in vicinity of the origin is linear in order
to eliminate the velocity-squared dissipative term in (2.17) when the state is
approaching the origin.
An alternative approach towards chatter reduction for the time-optimal
control of a double integrator system is proposed by Workman et al. (1987),
2.2. Time-optimal control 27
where k1 and k2 are the control gains, and α is the acceleration discount factor
such that 0 < α ≤ 1. Furthermore, using the saturation function as the finite
slope approximation to the signum function, the PTOS control law is given by
uptos = − sat k2 (x2 − fptos (x1 )) . (2.19)
1 2
s
k1 = , k2 = ,
yl αyl
with α and yl being the free parameters. Furthermore, comparing with the
characteristic equation of a second-order system
s2 + 2ζωn s + ωn2 = 0,
the natural frequency and the damping ratio of the closed-loop system in the
linear region near the origin are given by
s
1 1
s
ωn = , ζ= . (2.20)
yl 2α
10
PTOS
TOC
5
u = −1
dx/dt
−5
u = +1
−10
−25 −20 −15 −10 −5 0 5 10 15 20 25
x
0
u
−1
0 1 2 3 4 5 6 7 8 9 10
0
x
−5
TOC NTOC PTOS
−10
0 1 2 3 4 5 6 7 8 9 10
4
2
dx/dt
−2
0 1 2 3 4 5 6 7 8 9 10
Time [s]
NTOC better utilises the control capacity during the deceleration stage of a
set-point tracking manoeuver shown in Figure 2.8.
Additional performance improvement to PTOS algorithm, called the mod-
ified PTOS or MPTOS is proposed in Dhanda and Franklin (2009). This work
introduces extra flexibility in the choice of switching curve, given by
− k1 x1
if |x1 | ≤ yl ,
fmptos (x1 ) = k2
q
− sgn(x1 )
2α|x1 | − J if |x1 | > yl .
It can be shown that PTOS is a special case of MPTOS. From the continuity
constraints of the switching function, the gains are obtained as
2αζ 2 2α
s
k1 = , k2 = 2ζ 2
,
yl yl
with α, yl and ζ being the free parameters. Compare to PTOS, the addition
of the closed-loop damping coefficient ζ as a calibration parameter lead to a
more effective way of shaping the system response. The closed-loop natural
frequency in vicinity of the origin is given as
2αζ 2
s
ωn = .
yl
Note that both PTOS and MPTOS have the acceleration discount factor
α as a tuning parameter. While it is desirable to set α arbitrarily close to 1 to
make full use of the actuation capacity, it is typically assigned from 0.5 < α <
0.8 in practise to reduce overshoot. It is evident from (2.20) that a larger α
will yield a smaller damping, and consequently a larger overshoot. Salton et al.
(2012) proposed two solutions to remove this impediment. The first approach
involves modification of the control law in the linear region by the introduction
of a dynamic damping scheme. The second approach engages modification of
the switching curve by the replacement of α with an exponentially varying
factor scheduled by x1 .
Another approach that build on the PTOS algorithm but uses a sliding
mode controller in the linear region is proposed in (You and Lee, 2000) and
(You and Kwak, 2006). Yet another alternative is presented in (Zhou et al.,
2001) and (Zhang and Guo, 2000), whereby the switching profile in PTOS
is approximated using a third-order polynomial, and the feedback control is
30 Chapter 2. Literature review and research aims
From the survey in this subsection, it is evident that there are three main
ideas in eliminating chatter in time-optimal control while preserving high track-
ing performance. First, the abrupt change in control input across the switching
curve is replaced by a gradual transition. Furthermore, overshoot can be re-
duced by constructing a control law with high damping near the goal state.
Third, performance improvement can be achieved by reducing the acceleration
discount factor of the switching curve, for example by the introduction of a
scheduled acceleration discount factor, or approximating the switching curve
using a third order polynomial.
2500
R(s) + C(s) G1 (s) = s(s+1)(s2 +s+2500) Y (s)
−
di (t + pi ) = di (t) do (t + po ) = do (t)
+ +
+ 1 +
R(s) + C(s) G2 (s) = s(s+1) Y (s)
−
2.3.1 Background
It is well-known that vibration is commonly undesirable in many engineering
systems, as it hinders normal operation of machines, deters performance and
induces fatigue failure in mechanical components. One paradigm of solution
to suppress vibration is to employ actuators available in the system to cancel
the oscillation. This is exemplified by many successful implementations in a
broad range of systems, including flexible towers (Kar et al., 2000), magnetic
resonance imaging (MRI) systems (Roozen et al., 2008), nanopositioning stages
(Fleming, 2010) and hard disc drives (Semba and White, 2008).
From the system point of view, vibrations can be categorised into resonance
and forced vibration. Figure 2.9 shows two examples of systems susceptible to
vibration, where the top plant, G1 (s) has a resonance mode at 500 rad/s, while
the bottom plant G2 (s) endures external disturbances at the input and output.
Note that brake judder is categorised as a forced-vibration excited by BTV,
which is cyclical with respect to wheel spins (Jacobsson, 2001). Therefore, it
is the compensation of external periodic disturbances that will be the focus of
this section.
A periodic disturbance, d(·) cyclical in the domain x can be characterised
by
d(x + p) = d(x), ∀x.
The domain can be either time or spatial position, with respective examples
including the oscillation of offshore landing platforms due to sea waves (Mar-
coni et al., 2002) and force ripple in linear motors (Zhao and Tan, 2005).
32 Chapter 2. Literature review and research aims
Although not all periodic disturbance has an a priori known period, for exam-
ple, the sea waves that hit on the offshore landing platforms, there are many
disturbances that occur in rotational machineries where the period can be as-
sumed to be known or measurable, with examples including the eccentricity
in disc drives (Sacks et al., 1996), rotor vibration in helicopters (Arcara et al.,
2000) and mold-level fluctuation in steel castings (Manayathara et al., 1996).
Since the wheel rotational speed in modern vehicles are measurable and readily
available online, the period of oscillation for BTV can be calculated. In the
following, both the feedback and feedforward methods that mitigate periodic
disturbances with known or measurable period are reviewed.
Filtering
s2 + 2ζ1 ωs + ω 2
F (s) = ,
s2 + 2ζ2 ωs + ω 2
where ω is the centre frequency and ζ1 , ζ2 are the damping ratios with ζ1 > ζ2 .
The filter structure with parallel realisation augmented to a baseline controller
2.3. Active vibration attenuation 33
Filter
Baseline
do
F (s)
controller Plant
+ +
+ u +
r + C(s) G(s) y
−
Figure 2.10: Block diagram of a disturbance filter structure with parallel real-
isation added on a baseline servo system (Zheng et al., 2006).
Output regulation
ẇ = Sw. (2.22)
2.3. Active vibration attenuation 35
Exosystem
w e
Generalised Plant
u y
Controller
When both the states x and w are available for feedback, that is
x
y = ,
w
u = Kx x + Kw w. (2.23)
In the general case of any measurement y is available for feedback, then the
output feedback controller is considered, given by
ẋc = Ac xc + Bc y,
(2.24)
u = Cc xc + Dc y,
ΠS = AΠ + Bu Γ + Bw ,
(2.25)
0 = Ce Π + Deu Γ + Dew .
In many engineering systems, such as the HDD or the vehicle wheel, their
rotational speeds are time-varying and may result in periodic disturbances with
time-varying frequencies. This type of disturbance is called non-stationary
sinusoidal disturbance, and can be modelled using a linear time-varying (LTV)
exosystem:
ẇ = S(t)w. (2.26)
Note the difference between (2.26) and (2.22) is the time-varying nature of S
in the former case. Output regulation for LTV system has rarely been consid-
ered until the last decade in (Ichikawa and Katayama, 2006; Shim et al., 2006;
Zhang and Serrani, 2006; Shim et al., 2010). While most output regulation
results for the LTI system can be extended to the LTV system, one notable dif-
ference between the two is the introduction of a differential regulator equation
(Ichikawa and Katayama, 2006), given in the form of
Π̇ + ΠS = AΠ + Bu Γ + Bw ,
(2.27)
lim (Ce Π + Deu Γ + Dew ) = 0.
t→∞
Compare to the linear matrix equality form of (2.25), obtaining a solution for
the differential equation (2.27) is much more involving. Nevertheless, existing
algorithms are available, such as those provided in (Shim et al., 2006) to obtain
solution for certain type of minimum phase system, and in (Shim et al., 2010)
for certain type non-minimum phase system.
Further extension is proposed recently to the output regulation of linear
parameter-varying (LPV) systems, whereby the rejection of non-stationary si-
nusoidal disturbance is treated in (Köroǧlu and Scherer, 2011b) for measurable
frequencies and in (Köroǧlu and Scherer, 2011a) for uncertain frequencies. Sim-
ilar to the treatise on LTV systems, the necessary and sufficient condition for
the existence of a controller is governed by the differential regulator equation
(2.27). Furthermore, the controller is in the form of a LPV system, with a set
of gains scheduled by the frequency.
Although the derivation of a scheduled controller that achieves output reg-
ulation is investigated, the implementation of such controller has not been
extensively investigated. For practical reasons, there exists a strong motiva-
tion to investigate the discretisation of the LPV output regulation controller
and thus forms the basis for another research question to be addressed in this
thesis.
2.3. Active vibration attenuation 37
Repetitive Controller
+
Q(s)e−ps B(s) Generalised
do
+ Plant
+ +
+ + y
r + G(s)
−
e−ps
M (s) =
1 − e−ps
γ̇ = −gy$
(2.28)
uaf c = γ T $,
R θ1
× g dt ×
+
cos (ωt) cos (ωt)
R θ2 +
× g dt ×
di Generalised
sin (ωt) sin (ωt) Plant
+ +
+ + y
r + G(s)
−
rapidly changing due to tracking transient). It is this case where the AFC
demonstrates its advantage by simply freezing the parameter updates while
keeping the cancellation scheme in operation. Furthermore, in processes where
the compensator is frequently re-initialised, it is easy to store the parameters
values for initialisation at the next startup.
However, it is noticed that the AFC literatures mainly focus on the rejection
of input disturbances. While the complete cancellation of input disturbances
with time-varying frequencies can be achieved by scheduling the AFC scheme
with the frequency, Kinney and de Callafon (2011) indicate that the same ap-
proach is not sufficient for output additive disturbances when the frequency
is rapidly varying. This limitation hinders its implementation on the brake
judder compensation problem as the BTV is an additive output disturbance
with a varying frequency, which corresponds to the changing vehicle speed.
Extending the AFC for output disturbances with time-varying frequencies is
essential for its implementation in BTV compensation. Therefore, this pro-
vides the motivation for further investigation and forms the basis for the final
research question to be addressed in this thesis.
2.3.4 Summary
Rejection of periodic disturbance is a classical field that has been examined
using a variety of methods, including the approaches based on internal model
principle and learning-type compensation methods. While active vibration
attenuation is a mature area of research, active reduction of BTV using an
EMB is a novel idea that has not been investigated before in the literature.
40 Chapter 2. Literature review and research aims
To identify the best possible clamp force tracking trajectory in the face of
nonlinear dynamics and system constraints of the EMB, and to use it as a
benchmark for system performance. The established benchmark can be utilised
for subsequent clamp force controller developments to fully utilise the hardware
capability of the EMB. Experimental validations are required to demonstrate
that the proposed controller is practically implementable and to assess its
performance.
force control loop. By utilising the achieved bandwidth, brake torque varia-
tion compensation schemes are designed to mitigate the brake judder problem.
The overall structure of the thesis is depicted in Figure 2.14, while the specific
goals of each chapter are detailed in the following.
Chapter 2 reviews recent developments in the control of an automotive
electromechanical brake, existing solutions for brake judder attenuation and
various controller designs that will be further developed in the context of EMB
application. The review highlights the need to actively attenuate judder di-
rectly from the source. Furthermore, two topics on the control design warrant
further investigation: the development of a constrained time-optimal controller
and the oscillatory disturbance compensation schemes.
In Chapter 3, experimental setup used for this study is introduced. Utilising
the test rigs, EMB models with different levels of complexity are developed and
parameterised, and the brake judder phenomenon is examined.
In Chapter 4, a fast and accurate parameter identification procedure is pro-
posed to support quick quality control checks and rapid model-based controller
tunings post-production.
In Chapter 5, time-optimal tracking control of an EMB is first investigated
using Pontryagin’s minimum principle. To facilitate practical implementation,
robust near-time-optimal tracking control is developed. The performance of
this controller is experimentally validated against the current industry prac-
tise, which is the cascaded proportional-integral controller. Finally, the model
predictive controller is presented as the generalisation of preceding designs.
In Chapter 6, a linear parameter-varying torque variation compensator is
developed to attenuate brake torque variation (BTV) accountable for judder
development. Emulation design is investigated to aid practical implementation
of this scheme in a digital controller. Furthermore, experimental results of BTV
compensation are presented and discussed.
In Chapter 7, an adaptive feedforward torque variation compensator is pro-
posed to attenuate BTV. This design is shown experimentally to be effective
for BTV compensation under both fixed and varying wheel speed, and demon-
strates the advantage of having wheel position measurements that are typically
absent in production vehicles.
The last chapter summarises the major contributions of this thesis and sug-
gests further investigations that would further supplement the present results.
44 Chapter 2. Literature review and research aims
Introduction (Ch. 1)
Conclusions (Ch. 8)
45
46 Chapter 3. Experimental setup and model descriptions
1
A self-energising brake, such as the Electronic Wedge Brake (EWB) uses the actual
braking drag force to increase the force with which the brake is applied, where wedge-shape
mechanisms are typically employed in the design (Dietrich et al., 2001).
3.1. Experimental facilities 47
1 Planetary gears, 1 2 3 4 5 6 7
2 Motor stator,
3 Motor rotor,
4 Ball screw shaft,
5 Piston,
6 Pad,
7 Calliper.
sensor mounted about a toothed gear wheel on the motor rotor2 . Numer-
ical differentiation is utilised to calculate the instantaneous motor velocity
from the motor position measurements. The three phase motor currents are
evaluated from the DC link current measurements obtained using a DC link
shunt resistor. The three phase motor currents are used to evaluate the motor
quadrature current, which is then utilised to calculate the instantaneous motor
torque. Additionally, the clamp force is measured using a rosette strain gauge
mounted on the outside of the EMB piston sleeve. Lastly, communication with
the EMB is accomplished via the controller area network (CAN) bus interface.
1 2 3 4
Figure 3.4: Bench top experimental setup. The individual components are:
1. Data acquisition board; 2. PC; 3. EMB; 4. Brake calliper stand with load
cells.
three HBM C9B load cells used to ascertain the clamp force measurements ob-
tained from the internal sensor of the EMB. Not shown in the figure is a Delta
Elektronika SM 45-70 power supply that provides 42 V DC to the EMB.
Vector CANape, a software package with scripting capability is utilised to
update clamp force reference and log measurements. Furthermore, embedded
programming for the EMB is coded in Matlab/Simulink with Real-Time
Workshop Toolbox.
1 2 3 4 5 6
physical models of the EMB in the view of choosing a low-order model that
is adequate for simulation studies. This is followed by the development and
parameterisation of the simulation model.
The torque that drives the mechanical plant is provided by a PMSM, which
motion is governed by three-phase AC voltage supplied by the electrical system.
Detailed modelling of the PMSM circuitry can be found in (Lyshevski, 1999).
Line (2007) shows that the application of the direct-quadrature-zero (or dq0)
transformation on the three-phase circuit of an EMB reduces the three AC
quantities to two DC quantities, where simplified analysis can be carried out.
The equivalent DC circuit equations are given by
did
Lm = −Rm id + Lm iq θ̇r + vd ,
dt (3.2)
diq
Lm = −Rm iq − Lm id θ̇r − λm θ̇r + vq ,
dt
where id , iq are the currents in the d- and q-axis respectively, similarly vd , vq
are the voltages in the d- and q-axis respectively, Lm , Rm are the motor syn-
chronous inductance and winding resistance respectively, λm is the magnitude
of the flux linkage, and θr is the electrical angle. Furthermore, the motor
torque is given by
Tm = K t i q , (3.3)
dθm
= ωm ,
dt
dω
Σemb,sim : J m = −Tl (θm ) − Tf (θm , ωm ) + Kt iq , (3.6)
dt
di
Lm q = −Rm iq − Kb ωm + vq ,
dt
where ωm is the motor velocity. In order to attain a complete simulation
model, the characteristics of load torque Tl (·) and friction torque Tf (·) need
to be identified.
The load torque, Tl accounts for the stiffness of brake pads, calliper and gears,
and is described by
Tl (θm ) = N Fcl (θm ), (3.7)
where N is the gear ratio and Fcl (θm ) is the clamp force as a function of motor
position.
The lumped stiffness of the EMB calliper can be characterised by the
amount of motor angular displacement needed in order to generate a set
amount of clamp force. To determine the stiffness curve, the motor position
was incrementally varied stepwise from the contact position until the maximum
clamp force was reached, where the contact position is defined when the brake
pads and disc have just came into contact. The clamp force measurements are
plotted against the motor position in Figure 3.6. A third-order polynomial of
the motor position is fitted to the measured clamp force using the least squares
method. Since the polynomial visually overlaps with the measurements, this
demonstrates that the polynomial approximates the clamp force well under
the loading condition when the applied clamp force is increasing. With θm = 0
defined at the contact position, the stiffness characteristics is modelled using:
k1 θ 3 + k2 θ 2
if θm ≥ 0,
m m
Fcl (θm ) = (3.8)
0
if θm < 0.
4
x 10
3 Measurement (Apply)
Measurement (Release)
Model (Fcl = −0.3535θ3m + 35.19θ2m)
2.5
2
Clamp Force [N]
Apply
1.5
1
Release
0.5
0
0 5 10 15 20 25 30 35
Motor Position [rad]
Figure 3.6: Comparison between the clamp force and motor position measure-
ments with the fitted stiffness curve (3.8).
starting from the maximum clamp force to the contact position. The hystere-
sis effect can be observed from Figure 3.6. This indicates that the clamp force
is dependent on not only the motor position, but also the direction or rate of
loading (or unloading). To capture the effect of hysteresis, a rate-dependent
stiffness model is suggested by Jo et al. (2010). However, hysteresis is not
considered here as the modelling error due to hysteresis is less substantial
in practise compared to the effect due to variations in pad thickness or pad
temperature (Schwarz et al., 1998).
The friction model parameters were identified using three set of experi-
ments. The first set of experiment identifies the load-dependent static friction
torque, Ts , where a series of break-away tests was executed in air-gap, featuring
small increments of motor torque from stationary until motion is detected. The
motor torque that initiated the motion corresponded to the torque that was
required to overcome the stiction. This set of experiment was executed in the
air-gap to isolate the effect of clamp force. From 1000 observations, the stiction
torque resembles closely to a normally distributed probability density function
with mean 4.68 × 10−2 N·m and standard deviation 0.90 × 10−2 N·m, as shown
in Figure 3.7. In the absence of clamp force, the static friction torque solely
comprises of the load-independent component, given by Ts0 = 4.68×10−2 N·m.
The second set of experiment for identifying the friction model parameters
was the break-away tests executed with applied clamp force. Due to the pres-
ence of clamp force, the torque that started the motion corresponded to the
sum of load torque (3.7) and static friction torque, given by
Tm = Tl + Ts = N Fcl + Ts .
Experimental results are shown in Figure 3.8, where each data point represents
the mean of 20 repeated trials. The measured break-away torque increases with
clamp force, and can be well-approximated using the linear regression model
fitted using the least squares approach. Therefore, the load-dependent static
friction torque, Ts can modelled as a linear function of clamp force, in the form
56 Chapter 3. Experimental setup and model descriptions
200
Avg: 0.0468
Static Friction, Ts [Nm]
0.06
Std: 0.0090
Number of samples
150
0.04
100
0.02 50
Measurement
Mean
0 0
0 500 1000 0.02 0.04 0.06
Sample number Static Friction, Ts [Nm]
Figure 3.7: Static friction measurements from break-away tests in the air-gap,
where the mean is 0.0468 Nm
of
Ts (Fcl ) = Ts0 + µs Fcl , (3.11)
where Ts0 and µs are the load-independent static friction torque and the coeffi-
cient of friction respectively. Employing the y-interception of the linear regres-
sion, the load-independent static friction force is found to be 4.95 × 10−2 N·m.
Furthermore, using the slope of the linear regression and the gear ratio,
N = 2.63 × 10−5 m/rad, the load-dependent static friction coefficient is calcu-
lated as 3.53 N·m/N.
Note that the routine used in the break-away tests with applied clamp force
possesses several drawbacks. As one could expect, the process of incremental
increasing motor torque until motion is detected can be lengthy. In order to
speed up the process, larger torque increments were used, which could possibly
have resulted in a slightly larger load-independent static friction torque esti-
mate (4.95 × 10−2 N·m) compared to the previously obtained value from the
break-away tests in air-gap (4.68 × 10−2 N·m). Therefore, to resolve the loss of
accuracy in the estimation of Ts0 , the measurement data is refitted using the
least squares approach, however with Ts0 constrained to 4.68 × 10−2 N·m. The
constrained linear fit is depicted in Figure 3.8, which visually overlaps with
the linear regression obtained without constraints. From the slope of the con-
strained linear fit, the coefficient of friction is recalculated as µs = 3.55 N·m/N.
Additionally, the data point distribution along the clamp force axis is sparse
at high clamp force (> 18 kN), and measurement data for the maximum al-
lowable clamp force (30 kN) was not acquired. Since high clamp force leads
3.2. Modelling of an electromechanical brake 57
1.6
Measurement
Least Squares Linear Fit (y = 6.16e−05*x + 0.0495)
1.4 Constrained Linear Fit (y = 6.18e−05*x + 0.0468)
1.2
1
Torque [Nm]
0.8
0.6
0.4
0.2
0
0 0.5 1 1.5 2 2.5
Clamp Force [N] 4
x 10
Figure 3.8: Measurements from break-away tests with applied clamp force.
to large load torque and friction torque, a larger motor torque is required to
balance the motion. However, a prolonged high torque demand results in high
temperature that can potentially damage the power electronics. This prohib-
ited the execution of this test at high clamp force region. Therefore, the data
collected does not cover the whole operating envelope, and fitting a model us-
ing this data implicitly presume the extrapolation of data. This highlights the
need of a quicker identification procedure that can be performed throughout
the whole operating capability of the EMB.
The final parameters to be identified in the friction model were the
Coulomb friction, viscous friction coefficient and the characteristic velocity of
the Stribeck friction, represented by TC , Tv and ωs respectively. To achieve this,
the last set of experiment was executed by running the motor with a constant
velocity reference. In order to isolate the identification of Coulomb friction and
viscous friction coefficients, the tests were done in the air-gap. Due to a combi-
nation of cogging torque and torque ripple, as shown in Figure 3.9, the motor
torque required to maintain constant motor velocity varies sinusoidally. By us-
ing the average values of motor velocity and torque obtained from each test, a
58 Chapter 3. Experimental setup and model descriptions
0.5
0
0 2 4 6 8 10 12 14 16 18
0.08
Mean: 0.0377
0.06
Torque [Nm]
0.04
0.02
0
0 2 4 6 8 10 12 14 16 18
Time [s]
Figure 3.9: Example of experimentally obtained friction time series data for
motor velocity reference at 1 rad/s.
0.15
0.1
0.05
0
0 50 100 150 200 250
Motor Velocity [rad/s]
System constraints
The operating limitations presence in the plant shall be included in the simula-
tion model or addressed in the controller design. The clamp force is limited at
60 Chapter 3. Experimental setup and model descriptions
3.2.3 Summary
This section begins with a brief survey on high-order EMB models in the
view of selecting a suitable model fidelity used for controller designs. This is
followed by the development of a simulation model, presented in (3.6)–(3.11).
Subsequently, model parameter identification is executed, in which the routines
are summarised in Procedure 3.1.
3
While the design limit for clamp force is at 40 kN, the 30 kN limit is deliberately chosen
in this work to avoid overheating the motor during the initial development stage.
4
Some of the preliminary tests were performed with assistance of Caroline Jekelius, an
exchange student studying towards a Dipl.-Ing. at TU Munich, Germany.
3.2. Modelling of an electromechanical brake 61
Mechanical
Electrical
The EMB power transmission unit incorporates a planetary gear set and a
ball screw. Using the components’ geometry obtained from proprietary design
information, the gear ratio can be found. Furthermore, the lumped moment
of inertia about the motor’s axis of rotation is computed using the CAD data
of the EMB. The motor torque constant, armature inductance and armature
resistance are inferred from manufacturer’s data. Furthermore, the parameters
for the stiffness curve and friction model are identified using Step 2 and Step 3
of Procedure 3.1. The identified EMB parameters are tabulated in Table 3.1.
Note that the model parameters for stiffness and friction are estimated
separately using specifically designed control maneuvers that allow isolation of
a particular set of parameters. For example, the procedures are divided into
four different tests: the stiffness curve identification, two sets of break-away
tests and the constant velocity tests, where the first test identifies the stiffness
curve parameters, and the latter three tests identify the friction model param-
eters individually. Also notice that these experiments are run in steady state
conditions. While these routines have successfully characterised the stiffness
and friction properties of the EMB, they are lengthy and time consuming. For
instance, the break-away tests without clamping and with clamping each took
over four hours to complete, while the constant velocity tests required approx-
imately half an hour to run. Although human labour can be reduced consider-
ably by using computer scripts that automates the tests5 , it is envisaged that
a fast parameter identification procedure is required during production and
maintenance, where rapid quality control checks and model-based controller
tuning are conducted.
To check the accuracy of the estimated parameters and the simulation
model, the parameterised model for the mechanical plant is validated against
independently obtained experimental data. The measured motor torque is
used as the input for the simulation model to produce outputs in clamp force
and motor velocity. Comparison of experimental measurements and model
predictions is illustrated in Figure 3.12. The experimental and simulation
results are qualitatively matching, however, slight overshoots are observed in
the simulated responses. Note that the measured motor torque, which is also
the model input is contaminated with measurement noise. Since the input
measurement noise is not pre-filtered, it is not surprising to find the small
5
All tests are automated using scripts written in Vector CANape.
3.3. Development of a control-oriented EMB model 63
4
x 10
3
2
Fcl [N]
1 Experiment
Model
0
0 0.05 0.1 0.15 0.2 0.25 0.3
400
ωm [rad/s]
200
−200
0 0.05 0.1 0.15 0.2 0.25 0.3
4
Tm [Nm]
−2
0 0.05 0.1 0.15 0.2 0.25 0.3
Time [s]
Current Controller
1
s kci +
EMB Fcl
1 iq,r vq
Tm,r Kt +
Actuator
(Σemb,sim )
−
iq kcp +
Figure 3.13: Approximating the dynamics of the current control loop and the
EMB actuator using a control-oriented model.
3.3. Development of a control-oriented EMB model 65
dθm
tmech = tmech ωm ,
dt
dωr Rm
tmech =− (Tl + T̄f ) + ir ,
dt Kt Kb Ω
(3.16)
dir Rm + kcp kci kcp
telec =− ir − ωr + e0 + ir,r ,
dt Rm Kb Ω Rm
deo tmech Kb Ω
tmech = (ir,r − ir ),
dt Rm
where tmech = JRm /(Kt Kb ) is the mechanical time constant and telec = Lm /Rm
is the electrical time-constant. Using the model parameters given in Table 3.1,
tmech and telec are 1.01 × 10−2 and 1.13 × 10−3 respectively. Since tmech telec ,
tmech is considered as the time unit and the state equation (3.16) is rewrote
with respect to a dimensionless time variable τr = t/tmech as
dθm
= tmech Ωωr , (3.17a)
dτr
dωr Rm
=− (Tl + T̄f ) + ir , (3.17b)
dτr Kt Kb Ω
telec dir Rm + kcp kci kcp
=− ir − ωr + e0 + ir,r , (3.17c)
tmech dτr Rm Kb Ω Rm
deo tmech Kb Ω
= (ir,r − ir ). (3.17d)
dτr Rm
Equation (3.17) is in the standard form if the following dimensionless pa-
rameter is introduced:
telec Lm Kt Kb
ε1 = = 2
tmech JRm
Note that the dynamics of dir /dτr is fast when ε1 is small, therefore leading to
a rapid convergence of ir to the equilibrium of (3.17c). By setting ε1 to zero,
the equilibrium of (3.17c) is procured as
!
Rm kci kcp
ir = −ωr + e0 + ir,r . (3.18)
Rm + kcp Kb Ω Rm
3.3. Development of a control-oriented EMB model 67
tcc deo Kb Ω
= −e0 + (ωr + ir,r ), (3.19)
tmech dτr kci
where tcc = (Rm + kcp )/kci is the current controller time constant. Using the
Rm value provided in Table 4.3, the proportional gain kcp = 0.3385 and the
integral gains kci = 199, tcc can be calculated as 1.95 × 10−3 , thus tcc tmech .
Introducing a dimensionless parameter
which can be made small by proper selection of kcp and kci , the equilibrium of
(3.19) can be computed as
Kb Ω
e0 = (ωr + ir,r ). (3.20)
kci
ir = ir,r .
iq = iq,r , (3.21)
where xd (k) = [θm (k), ωm (k)]T is the state vector at time step k and x0 is
a given initial condition of the state vector. Employing the forward-Euler
method, fd (·) can be calculated as
fd = xd (k) + ts f xd (k), Tm (k), θ , (3.25)
where ts is the step size. The function f (·) and h(·) are given in (3.23).
Alternative discretisation methods, such as the family of Runge-Kutta methods
can also be used to derive alternative forms of (3.25).
Note that three EMB models have been presented from Section 3.2 to Sec-
tion 3.3, namely the simulation model Σemb,sim in (3.6), the control-oriented
model Σemb in (3.22), and the discretised model Σemb,d in (3.24). The step
responses and sinusoidal responses of these three models are compared in Fig-
ure 3.14. The responses of the three models visually overlap with each other.
4
x 10
4
Full−order nonlinear model
Fcl [N]
Reduced−order model
2 Discrete model
0
0 0.05 0.1 0.15 0.2 0.25 0.3
400
30 kN Step
ωm [rad/s]
200 20 kN Step
0
10 kN Step
−200
0 0.05 0.1 0.15 0.2 0.25 0.3
5
Tm,r [Nm]
0
10 kN Step 30 kN Step
20 kN Step
−5
0 0.05 0.1 0.15 0.2 0.25 0.3
Time [s]
4
x 10
4
Full−order nonlinear Reduced−order Discrete
Fcl [N]
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
400
ωm [rad/s]
200
−200
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
5
Tm,r [Nm]
−5
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Time [s]
(b) Sinusoidal responses. Grey lines denote responses for 10 + sin (20π) kN and
black lines denote responses for 20 + sin (20π) kN.
Figure 3.15: Setup for disc rotor runout measurements. The individual parts
are: 1. Brake disc rotor with runout and thickness variation; 2. Dial indicator.
machining, uneven wear, uneven corrosion and uneven friction film generation.
Typically, these irregularities occur fewer than five times per rotor revolution,
with the most threatening case occurs in rotor with one or two blemishes.
This thesis deals with the compensation of low-order BTV that causes
the most common type of brake judder, which is called cold judder. The
cold judder typically occurs when the brake temperature is less than 100◦ C.
Additionally, the low-order BTV is due to the uneven wear of disc rotor, where
the level of evenness can be determined from runout and thickness variation
(DTV). For most manufacturers, the manufacturing tolerances for runout and
DTV in a disc rotor are less than 80 µm and 10 µm respectively (de Varies
and Wagner, 1992).
The disc rotor employed for experiments in this work is illustrated in Fig-
ure 3.15 and was obtained from an in-service vehicle with brake judder. Using
a dial gauge, runout for both surfaces of the rotor was measured, where the
measurements are shown in Figure 3.16. The maximum runout for the inner
surface and the outer surface are 87 µm and 97 µm respectively, demonstrating
higher values than the designed tolerance.
DTV is computed from the difference in runout of both surfaces, which is
3.4. Modelling of the characteristics of brake torque variation 71
0.1
−0.05
−0.1
0 50 100 150 200 250 300 360
0.1
Outer face runout [mm]
0.05
−0.05
−0.1
0 50 100 150 200 250 300 360
Disc Rotation [deg.]
Figure 3.16: Runout measurements for inner and outer faces of the brake disc.
k=1
where the amplitude of the k-th harmonic is ad,k and the phase-shift is φd,k .
The components of different frequencies in the DTV waveform can be de-
termined using the fast Fourier transform (FFT) method. Employing the fft
function in Matlab, the Fourier coefficients are computed, which are depicted
in Figure 3.17. It is evident that the first harmonic has the largest component,
followed by smaller amplitudes of the second and the third harmonics. Util-
ising the Fourier coefficients, various orders of approximation can be made.
For example, the m-th order of approximation is computed using (3.26) up to
72 Chapter 3. Experimental setup and model descriptions
(a) 0.02
0.015
0.01
0.005
DTV [mm]
−0.005
Measurement
−0.01 1st approx.
−0.015 2nd approx.
3rd approx.
−0.02
0 50 100 150 200 250 300 360
(b) Disc Rotation [deg.]
0.015
Amplitude [mm]
0.01
0.005
0
0 1 2 3 4 5 6
Harmonics
Figure 3.17: (a) Disc thickness variation measurements versus the first three or-
ders of approximation; (b) Fourier transformation of the measurements shows
that the signal is dominated by the first three harmonics.
where 2µrd Fcl represents the mean brake torque, and Tbtv (θd ) represents the
variation in brake torque induced by rotor geometrical irregularities. In order
to examine the variations of brake torque, the disc rotor with runout and DTV
was rotated at one hertz with 750 kN of clamp force applied. The brake torque
measurements are shown in Figure 3.18, demonstrating a periodic oscillation
in brake torque. Note that the brake torque can be approximated using an
n-order Fourier series, given by
n
Tb = ab,0 + ab,k cos (kθd + φb,k ), (3.28)
X
k=1
(a) 160
Mea. DC 1st. 2nd. 3rd.
150
Brake Torque [Nm]
140
130
120
110
100
90
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
(b) Disc Rotations
15
Amplitude [Nm]
10
0
0 1 2 3 4 5 6
Harmonics
Figure 3.18: (a) Brake torque variation measurement versus the mean and
the first three orders of approximations; (b) Fourier transformation of the
measurements shows that the signal is dominated by the mean and the first
three harmonics.
k=1
Similar to the analysis conducted on the DTV waveform, the FFT is applied
to the brake torque measurements to calculate the components of different
frequencies. Apart from the mean, which has the largest amplitude given by
ab,0 = 112.85 Nm, the first harmonic has the second largest component at ap-
proximately 15 Nm. This is followed by smaller components of the second-order
and the third-order harmonics. The second-order and the third-order approxi-
mations are very close and visually coincide, and they match the measurements
well. Furthermore, utilising the effective rotor radius of rd = 127 mm and the
mean of measured clamp force, Fcl = 732.28 N, the effective coefficient of
friction can be calculated as µ = 0.607.
The measurements of the DTV and BTV clearly indicate that they are peri-
odic with respect to disc rotor rotations. Therefore, the periodic BTV function
74 Chapter 3. Experimental setup and model descriptions
established in (3.29) will be used in the latter chapters for compensators syn-
thesis. However, the amplitude and phase-shift of the BTV are unknown a
priori but are assumed to be time-invariant, and they have to be adapted by
the proposed compensators. In practice, these quantities may vary slowly rel-
ative to the disc rotations. Since both of the DTV waveform and the BTV
measurements can be well approximated by their respective second-order ap-
proximations, therefore BTV compensators of up to two orders are sufficient
in implementation.
3.5 Conclusions
An overview of the experimental facilities that are used in this research are
presented in this chapter. Subsequently, the modelling work is started in Sec-
tion 3.2 with a review on high-order models of EMB in the interest of deciding
a suitable model fidelity for control performance simulations. An EMB model
used for simulation is then presented in (3.6), which comprises of the torque
equation evaluated around the motor rotational axis and the electrical circuit
dynamics for quadrature current. The overall nonlinear stiffness characteristics
and nonlinear friction of the EMB calliper are included in the model. A model
identification routine summarised in Procedure 3.1 is presented to determine
the characteristics of stiffness and friction, and also to estimate the parame-
ters contained in these models. The routine is designed to run in steady-state
conditions, and therefore it is lengthy and time-consuming. It demanded nine
hours of data collection even with the scripted automatic testings, plus addi-
tional hours for post-processing of data. Although this routine has successfully
parameterised the stiffness model and the friction model, it is not applicable
for post-production quality control checks and quick model-based controller
tunings.
To facilitate controller developments, a reduced-order nonlinear control-
oriented model for the EMB is presented in Section 3.3. The proposed control-
oriented model incorporates the dynamics of the current controller and the
dynamics of the EMB actuator. Since the dynamics of the current controller
and the dynamics of the electrical circuits in the EMB actuator are relatively
fast compared to the mechanical components, singular perturbation technique
is employed to justify the reduction of the current controller and electrical
3.5. Conclusions 75
+
EMB Fcl Brake Disc +
1 Current vq Actuator with Nominal Tb
Tm,r Kt + Controller
Σemb,sim , Thickness
− (3.6) in
iq Sec. 3.2
ωm
Fcl
Tb
Figure 3.19: A block diagram illustrating the models that contribute to brake
torque generation.
Rapid electromechanical
brake model parameter
identification
In order to support quick quality control checks and rapid model-based con-
troller tunings post-production, a fast and accurate parameter identification
procedure is desirable. However, the parameter identification routine sum-
marised in Procedure 3.1 is lengthy and time consuming, where it required
nine hours of data collection even with automated testings, in addition to
extra hours for post-processing of data. Additionally, measurement noise is
inevitable in parameter identification experiments, which leads to estimation
inaccuracies. These considerations motivate the development of a rapid model
parameter identification procedure for an electromechanical brake (EMB) that
incorporates noise models.
The design of the rapid parameter identification procedure is commenced in
two stages. First, the optimal experimental design is conducted in Section 4.1.
This involves planning an optimal output trajectory to minimise the time re-
quired for experiments needed to infer model parameters from input-output
measurements. This trajectory has to be achievable by complying to the input
and output constraints. In the second stage, parameter estimation algorithms
are devised in Section 4.2 to identify model parameters from the input-output
measurements. Two estimation methods, namely the output error method
77
78 Chapter 4. Rapid EMB model parameter identification
and the prediction error method are formulated. Their accuracy and ease of
implementation are also compared and discussed. Finally, the parameterised
EMB model is validated against independently obtained experimental mea-
surements.
where xd (k) := [θm (k), ωm (k)]T is the state vector at time step k and x0 is
a given initial condition of the state vector. The functions fd (·) and h(·) are
given in (3.25) and (3.23) respectively. The matrix, Bv associated with the
process noise is given by
0
Bv = 1 . (4.2)
J
The plant dynamics are affected by the process noise, vi (k) and the associated
matrix Bv , while the output measurement ym (k) is corrupted by the output
measurement noise, vo (k). Note that the EMB parameters,
h i h i
θ = θ1 θ2 · · · θ7 = k1 k2 TC Ts0 Tv µs ωs .
4.1. Design of an optimal trajectory for parameter identification 79
where E{·} denotes the expected value. By exploiting (4.3), an estimate for
the variance of the parameter estimation error can be calculated from (Raol
et al., 2004):
1X n T
R= ym (k) − yd (k) ym (k) − yd (k) . (4.4)
n k=1
To obtain ym (·) and yd (·), an effort can be made to drive the plant using
an arbitrary chosen input trajectory {Tm (k), k = 1, . . . , n}, where the output
ym (k) are measured. Then, the model output, yd (·) can be computed from
(4.1) using the input sequence.
The task of parameter identification is to estimate θ from the knowledge
of the input and output sequences, which are denoted by U = {Tm (k), k =
1, . . . , n} and Y = {ym (k), k = 1, . . . , n} respectively. Since the true parame-
ters, θ∗ are unknown in real-life scenario, the exact parameter estimation errors
cannot be obtained. Alternatively, the relative accuracy of the estimation is
evaluated from the bias and the covariance of the estimated parameter. An
unbiased estimator with maximum likelihood is often assumed during the ex-
perimental design phase so that the it can be accomplished independently of
the estimator used (Mehra, 1974). Note that the covariance of any unbiased
estimator is bounded by the Cramér-Rao lower bound, given by
n o
E (θ − θ∗ )(θ − θ∗ )T ≥ M −1 , (4.5)
and p(Y |U, θ) denotes the conditional probability density, also known as the
likelihood function. Furthermore, the lower bound of the covariance can be
achieved by utilising the maximum likelihood estimator, which maximises the
conditional probability density:
Assumption 4.1. The Fisher information matrix for the input output pairings
in the trajectory, M is nonsingular.
Remark 4.1. Note that a nonsingular M is necessary for the existence of the
Cramér-Rao lower bound matrix, and can be provided through appropriate
design of the experimental trajectory.
The experimental design can be approached by optimising the input se-
quence, U such that the resultant output trajectory yields a small covariance
in the estimated parameters. As established by the Cramér-Rao lower bound
(4.5), the lower bound of the maximum likelihood estimator covariance is given
by the inverse of the Fisher Information Matrix, M −1 . Therefore, a scalar
measure of M −1 can be used as a performance metric for optimising the tra-
jectory. To this end, the following scalar measures of performance based on M
are commonly considered (Mehra, 1974):
The D-optimal design has a notable advantage over the A-optimal design, in
which it is invariant under scale changes in the parameters and linear trans-
formations of the output (Mehra, 1974). This attractive feature has motivated
the adoption of D-optimal design in this work.
Note that the EMB model have magnitudes differing by factors of up to
106 . To facilitate meaningful comparison between variances and to avoid nu-
merical sensitivity during optimisation, parameters and outputs are scaled.
The normalised parameter vector, θ̄ = [θ̄1 , θ̄2 , . . . , θ̄7 ]T is introduced, which is
defined by
θi
θ̄i = , i = 1, 2, . . . , 7 (4.8)
θnom,i
4.1. Design of an optimal trajectory for parameter identification 81
The scaled outputs h̄ have reduced ranges with the same order of magnitude,
n o
given by h̄ := [h̄1 , h̄2 ]T : h̄1 ∈ [0, 3] .
Furthermore, the scalar performance measure of the D-optimal design can
be converted into a convex function using log det (M −1 ) (Boyd and Vanden-
berghe, 2004). It can be shown that M is a positive semidefinite symmetric
matrix. If M is nonsingular, then − log det (M ) = log det (M −1 ) can be em-
ployed in order to avoid numerical issues during the matrix inversion when M
has a bad condition number.
n
!T !
dh̄ dh̄
M= −1
k = 1, 2, . . . , n (4.11a)
X
R ,
k=1 dθ̄ k dθ̄ k
!
dh̄ h i
= dh̄
dθ̄1
dh̄
dθ̄2
··· dh̄
dθ̄7
, (4.11b)
dθ̄ k
! ! !
dh̄ ∂ h̄ ∂ h̄
= xθ̄i ,k + , i = 1, 2, . . . , 7 (4.11c)
dθ̄i k
∂xTd k ∂ θ̄i k
! !
∂fd ∂fd
xθ̄i ,k+1 = x + , xθ̄i ,1 = 0. (4.11d)
∂xTd k θ̄i ,k ∂ θ̄i k
The arguments of fd (·) and h(·) are omitted and the time steps are written in
the subscripts for clarity in presentation. For completeness, the expressions in
(4.11) are expanded in Appendix B.
4
x 10
3
2
Fcl [N] 1
0.1 0.2 0.3 0.4
0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
400
ωm [rad/s]
200
0
−200
−400
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
4
2
Tm [Nm]
0
−2
−4
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
Time [s]
presented. It is observed that the values of both criteria1 decrease with the
experimental length, implying a longer experiment provides better parameter
estimation accuracy. This is because a longer experiment allows more time for
full system dynamics to be excited and more data to be collected. However, a
decreased rate of improvement is observed, where the reduction in the perfor-
mance criteria is more significant for experiments shorter than 0.18 seconds.
This is because for shorter experiments, the maximum values of clamp force
transients increase significantly with a slight increase in experimental length.
However, for experimental length longer than 0.18 seconds, the maximum val-
ues of clamp force trajectories remain at approximately 23 kN, as shown in
Figure 4.1. To strive a balance between the estimation accuracy and the time
required for experiments, 0.3 seconds is chosen as the length of experiment,
and the corresponding optimal trajectory will be applied in the next subsec-
tion for parameter estimation. While the experimental length of 0.2 seconds
is adequate, a slightly longer experiment is chosen to account for situations
1
A few points for tr (M −1 ) are missing due to numerical problem faced during the
evaluation of M −1 as M has bad condition number.
84 Chapter 4. Rapid EMB model parameter identification
10 800
−log det (M)
0 tr (M−1) 700
−10 600
−20 500
−log det (M)
tr (M−1)
−30 400
−40 300
−50 200
−60 100
−70 0
0.1 0.15 0.2 0.25 0.3 0.35 0.4
Length of Experiment [s]
where softer brake pads are used and longer rise-times are expected compared
to the nominal case.
Since the proposed rapid parameter identification procedure is intended for
quick post-production quality control checks and model-based controller turn-
ing during maintenance, it is therefore presumed that a feedback controller,
possibly with unknown structure or gains has already been implemented. To
reduce the risk of damaging the hardware, the parameter identification is ex-
ecuted with feedback, where the identified optimal clamp force trajectory is
employed as a reference trajectory and tracking is handled by the existing feed-
back controller. The experimentally obtained input and output measurements
will then be used to infer the model parameters in the following subsection.
Procedure 4.1 (Parameter estimation using output error method). Given the
plant model (4.1) and a sequence of measured input U := [Tm (1), . . . , Tm (n)]T
and measured output Y := [ym (1), . . . , ym (n)]T , the OEM finds the model
parameters θ and the initial condition x0 in order to minimise the error between
the measured and the model outputs, which can be formulated as the following
(Raol et al., 2004):
1X n T n
min ym (k) − yd (k) R−1 ym (k) − yd (k) + log det (R),
x0 ,θ 2 k=1 2
subject to xd (1) = x0 ,
xd (k + 1) = fd xd (k), Tm (k), θ , k = 1, . . . , n
yd (k) = h xd (k), θ ,
1X n T
R= ym (k) − yd (k) ym (k) − yd (k) .
n k=1
(4.12)
where A(k − 1) is the state transition matrix. These predictions are subse-
quently corrected in the update phase using measurements at k. This involves
the calculation of the Kalman gain, given by
The states and covariance are then updated using the following expressions
with the Kalman gain:
h i
x̂(k|k) = x̂(k|k − 1) + K(k) ym (k) − h(x̂d (k|k − 1), θ) ,
h i (4.16)
P (k|k) = I − K(k)C(k) P (k|k − 1).
The state transition matrix A(k − 1) and the observation matrix C(k) are
defined to be the following Jacobians:
∂fd ∂h
A(k − 1) = , C(k) = , (4.17)
∂xTd x̂(k−1|k−1),Tm (k−1)
∂xTd x̂(k|k−1)
4
x 10
3
Non−optimal experimental design
2
Fcl [N]
0
0 0.05 0.1 0.15 0.2 0.25 0.3
400
ωm [rad/s]
200
−200
0 0.05 0.1 0.15 0.2 0.25 0.3
4
Tm [Nm]
−2
0 0.05 0.1 0.15 0.2 0.25 0.3
Time [s]
Table 4.1: Mean and variance of the normalised estimated parameters obtained
using the output error method (OEM) and the prediction error method (PEM)
with the optimal experimental design.
Normalised OEM PEM
Parameter Mean Variance Mean Variance
k1 /k1,nom 0.7985 0.004162 0.8438 0.004331
k2 /k2,nom 0.9015 0.003235 0.9363 0.000898
TC /TC,nom 1.6624 0.014324 1.6682 0.006034
Ts0 /Ts0,nom 0.9298 0.014876 0.9212 0.008480
Tv /Tv,nom 2.2158 0.059713 2.4498 0.065986
µs /µs,nom 1.0508 0.018334 0.8747 0.002355
ωs /ωs,nom 0.4261 0.012167 0.4063 0.026712
1.25 1.9
1.2 Est. Parameter
Mean 1.8
TC / TC,nom
1.7
1
1.6
0.9
1.5
0.8
0.75 1.4
0.5 1 1.5 1.8 2 2.2 2.4 2.6 2.8
k1 / k1,nom Tv / Tv,nom
1.25 0.5
1.2
0.4
ωs / ωs,nom
1.1
µs / µs,nom
0.3
1
0.2
0.9
0.1
0.8
0.75 0
0.5 1 1.5 1.8 2 2.2 2.4 2.6 2.8
Ts0 / Ts0,nom Tv / Tv,nom
Figure 4.4: The normalised estimated model parameters obtained using the
output error method.
90 Chapter 4. Rapid EMB model parameter identification
1.25 1.9
1.2 Est. Parameter
Mean 1.8
TC / TC,nom
1.1 50% Conf. Region
k2 / k2,nom
1.7
1
1.6
0.9
1.5
0.8
0.75 1.4
0.5 1 1.5 1.8 2 2.2 2.4 2.6 2.8
k1 / k1,nom Tv / Tv,nom
1.25 0.5
1.2
0.4
ωs / ωs,nom
1.1
µs / µs,nom
0.3
1
0.2
0.9
0.1
0.8
0.75 0
0.5 1 1.5 1.8 2 2.2 2.4 2.6 2.8
Ts0 / Ts0,nom Tv / Tv,nom
Figure 4.5: The normalised estimated model parameters obtained using the
prediction error method.
using the newly acquired optimised parameters. It is evident that the simulated
responses obtained using the optimised parameters match the experimental
results better. This result indicates that the optimised parameters are more
accurate than the nominal values.
The performance of estimation can be compared using the variance of the
estimated parameters. The sum of variances given by the output error method
and the prediction error method are 0.0048024 and 0.0038791 respectively.
Since the sum of variances given by the prediction error method is smaller,
this suggests that it renders better estimations. This can also be examined by
comparing the 50% confidence region depicted in Figure 4.4 and Figure 4.5,
where the estimations provided by the prediction error method have smaller
50% confidence regions. This improvement is resulted from the consideration
of process noise in the prediction error method, where the covariance Q is
estimated to be 0.00020154, suggesting that the motor torque measurements
have errors with standard deviation of 0.014196 Nm. Note that the estimated
standard deviation in torque measurement error is relatively substantial com-
4.2. Parameter estimation 91
4
x 10
6
Exp. Nom. OEM PEM Non−opt.
Fcl [N]
4
0
0 0.05 0.1 0.15 0.2 0.25 0.3
400
ωm [rad/s]
200
−200
0 0.05 0.1 0.15 0.2 0.25 0.3
4
Tm [Nm]
−2
0 0.05 0.1 0.15 0.2 0.25 0.3
Time [s]
pared to the friction torque at very slow speed, as it has the same order of
magnitude compared to the break-away friction torque.
Further model validation is performed using experiment with sinusoidal
trajectory, as shown in Figure 4.7. There are some disagreements in the simu-
lated and the experimental responses. A closer look at the velocity responses
indicates that the model shows lockups when the velocity crosses zero while
the clamp force transient is at its peak. However, the velocity lockups are not
observed in the experimental measurements. This suggests that the response
is very sensitive at very low speed, as a small error in motor torque measure-
ment or minor estimation error in friction characteristics at very low speed
may result in velocity lookups, which then lead to significant deviation in the
clamp force trajectory. Nevertheless, in the period where velocity lookups do
not happen, for example in between 1.5 seconds to 1.8 seconds, the simulated
responses are qualitatively matching the measurements well.
The correction mechanism provided by the extended Kalman filter is signif-
icant for model response sensitive to the input measurement noise. Figure 4.8
92 Chapter 4. Rapid EMB model parameter identification
4
x 10
3
Experiment OEM PEM
2
Fcl [N]
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
200
ωm [rad/s]
−200
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
2
Tm [Nm]
−1
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Time [s]
shows the mean and covariance estimates provided by the extended Kalman
filter for the sinusoidal trajectory. In the clamp force response, the mean and
the region bounded by the variance are visually overlapping, implying a good
estimate is obtained. On the other hand, the variance of the estimated velocity
increases significantly at zero-crossings. However, due to the correction fea-
ture, lookup is not noticeable in the estimated velocity mean. Therefore, the
parameter estimation provided by the prediction error method is more robust
against input measurement noise. On the other hand, due to the inclusion of
extended Kalman filter, the prediction error method requires ten times more
computational time compared to output error method in this application.
Figure 4.8: Kalman estimates of mean and variance for sinusoidal tracking
response.
Table 4.2: Mean and variance of the normalised estimated parameters obtained
using the output error method with the non-optimal experimental design.
Normalised Mean Variance
Parameter
k1 /k1,nom 0.8147 0.004884
k2 /k2,nom 0.8912 0.004399
TC /TC,nom 1.6093 0.009592
Ts0 /Ts0,nom 0.9251 0.019993
Tv /Tv,nom 1.6597 0.018759
µs /µs,nom 0.9616 0.021593
ωs /ωs,nom 0.3837 0.013261
the estimated parameters and plotted in Figure 4.6. Compared to the experi-
mental measurements, an overshoot of velocity is observed. Additionally, the
trajectories of clamp force and velocity deviate from the measurements after
0.2 seconds, hinting an underestimation of friction torque. The sum of vari-
ances of the estimated parameters is 0.0072916, which is significantly higher
compared to the optimal experimental design.
94 Chapter 4. Rapid EMB model parameter identification
1.25 1.9
1.2 Est. Parameter
Mean 1.8
TC / TC,nom
1.1 50% Conf. Region
k2 / k2,nom
1.7
1
1.6
0.9
1.5
0.8
0.75 1.4
0.5 1 1.5 1.4 1.6 1.8 2 2.2 2.4
k1 / k1,nom Tv / Tv,nom
1.25 0.5
1.2
0.4
ωs / ωs,nom
1.1
µs / µs,nom
1
0.3
0.9
0.8 0.2
0.75
0.5 1 1.5 1.4 1.6 1.8 2 2.2 2.4
Ts0 / Ts0,nom Tv / Tv,nom
Figure 4.9: The normalised estimated model parameters obtained using a non-
optimal experimental design and the output error method.
Finally, the optimised model parameters provided in Table 4.3 will be used
in the following chapters for simulations and controller designs.
4.3. Conclusions 95
4.3 Conclusions
A rapid parameter identification procedure for the prototype automotive elec-
tromechanical brake (EMB) is proposed in this chapter, which is intended to
support post-production quality control checks and quick model-based con-
troller tunings. The proposed procedure parameterised the EMB model in-
troduced in Chapter 3. Compared to the incumbent approach summarised in
Procedure 3.1, which uses a few different experimental manoeuvres that isolate
the identification of certain sets of model parameters, the rapid approach uses
a single optimised experimental trajectory, therefore greatly reduces the time
required for parameter identification. Specific outcomes of this work include:
tal design has reduced the time required for experiments from nine hours
to one minute.
• The parameter estimation problem is formulated using both output er-
ror method and prediction error method to infer model parameters from
experimental measurements. Both methods explicitly considers the out-
put measurement noise, but the prediction error method also takes into
account of the process noise. It is observed that the prediction error
method provides a marginally better estimation accuracy compared to
the output error method but requires significantly longer evaluation time.
This may be an important consideration for practical implementation of
the EMB model parameter estimation.
• The parameterised model is validated against independently obtained ex-
perimental data. Compared to the parameters estimated in Section 3.2,
the simulated responses of the newly estimated parameters provide a
better match with real measurements. This indicates that the proposed
rapid parameter identification procedure is able to obtain more accurate
parameter estimations in shorter time. The optimised model parameters
are listed in Table 4.3 and they will be used for controller development
in subsequent chapters.
5
97
98 Chapter 5. High performance clamp force tracking control
where
ωm
f (x, Tm ) = 1 . (5.2)
J
−N F (θ
cl m ) − T̄ (θ
f m , ωm ) + T m
The stiffness characteristics Fcl (·) and friction model T̄f (·) are given by (3.8)
and (3.12) respectively. The state vector consists of the motor position and ve-
locity, x := [θm , ωm ]T ∈ X and the input comprises the motor torque, Tm ∈ U,
where U is compact. Furthermore, f : X × U → R2 with f (0, 0) = [0, 0]T .
Note that the regions of X and U are dictated by the state and input con-
straints. The motor position is limited by the clamp force range. Additionally,
5.1. Time-optimal tracking control 99
the velocity saturates due to the maximum voltage and the maximum rotary-
current-frequency of the servo amplifier output. Therefore, the state region is
given by
n o
X = x = (θm , ωm ) ∈ R2 : Fcl (θm ) ∈ [0, Fcl max ], |ωm | ≤ ωm max . (5.3)
The control problem dealt with is the clamp force feedback control. It is
studied under the state feedback control framework. However, since only the
velocity and clamp force are available for feedback, the measured clamp force
is used to deduce the motor position. By utilising the one-to-one relationship
of the clamp force function to the motor position, the clamp force reference
and clamp force measurement are represented by the equivalent motor position
reference and measurement respectively. The corresponding state reference is
given by xr = (θm,r , ωm,r ).
In order to deliver demanding clamp force requests from higher level con-
trollers, such as the ABS and ESC during safety-critical operations, the clamp
force controller is expected to have quick tracking performance. This moti-
vates the utilisation of the time-optimal control, as it is expected to provide
the quickest possible tracking response.
subject to (5.1)–(5.2),
Tm (t) ∈ U, (5.5)
x(t) ∈ X ,
x(t0 ) = x0 ,
x(tf ) = xr (tf ).
100 Chapter 5. High performance clamp force tracking control
The state constraint X and control constraint U are defined by (5.3) and (5.4)
respectively.
where the dynamics of the states, θm and ωm are expressed in (5.1) and (5.2).
Furthermore, the Lagrangian L is defined by
L x(t), u(t), p(t) = H x(t), u(t), p(t)
+ µu1 (t)fcu1 Tm (t) + µu2 (t)fcu2 Tm (t)
+ µx1 (t)fcx1 ωm (t) + µx2 (t)fcx2 ωm (t) ,
where the multipliers µu1 , µu2 , µx1 , µx2 are non-negative. The constraints on
the motor torque, fcu1 , fcu2 can be expressed as
fcu1 Tm = Tm − Tm max ≤ 0,
(5.7)
fcu2 Tm = −Tm − Tm max ≤ 0,
∂L
ṗ1 = −
∂θm
µs tanh (εωm )
!
p2
= 3k1 θm
2
+ 2k2 θm N + , (5.9a)
J 1 + (ωm /ωs )2
∂L
ṗ2 = −
∂ωm
5.1. Time-optimal tracking control 101
" !
p2 Ts − TC
= −p1 + Tv + ε TC + 1 − tanh (εω m ) 2
J 1 + (ωm /ωs )2
2ωm (Ts − TC ) tanh (εωm )
#
− − µx1 + µx2 . (5.9b)
ωs2 (1 + ωm
2 /ω 2 )2
s
Proof. By substituting (5.1) and (5.2) into the Hamiltonian (5.6), proceeded
by the application of PMP (2.4), the following expression can be obtained:
p∗2
1 + p∗1 ωm
∗
+ −N Fcl (θm
∗
) − T̄f (θm
∗ ∗
, ωm ) + Tm∗
J
p∗
≤ 1 + p∗1 ωm + 2 −N Fcl (θm ∗
) − T̄f (θm
∗ ∗
, ωm ) + Tm , (5.10)
J
where the superscript ∗ denotes the optimal trajectory.
When p∗2 (t) > 0, the optimal control has to be Tm∗ (t) = −Tm max for (5.10)
to hold for all Tm (t) ∈ U. Conversely, when p∗2 (t) < 0, Tm∗ (t) = Tm max . Hence,
the time-optimal control input is given by
Tm∗ (t) = −Tm max sgn p∗2 (t) . (5.11)
Proposition 5.2 (Optimal trajectories which lie on the boundary of the state
constraint). Consider the time-optimal tracking problem (5.5). Assume that
the velocity constraint is reached, ωm (t) = ±ωm max and the optimal costate
p∗2 (t) = 0, for t ∈ [t1 , t2 ], t0 ≤ t1 < t2 ≤ tf . If Tm∗ (t) is the time-optimal control,
it is necessary that Tm∗ (t) = N Fcl θm ∗
(t) + T̄f θm ∗
(t), ωm
∗
(t) for t ∈ [t1 , t2 ].
ṗ∗1 (t) = 0.
When negative motor velocity constraint is reached, ωm = −ωm max , the proof
can be repeated to obtain the same time-optimal control solution (5.15).
It has been shown in Proposition 5.1 and Proposition 5.2 that the time-
optimal trajectory consists of maximum acceleration and/or deceleration, and
possibly an intermediate singular arc with constant maximum velocity. This
demonstrates that the available motor torque capacity is fully utilised, and the
motor velocity is pushed to the limit in order to achieve a rapid response.
5.1. Time-optimal tracking control 103
1500 −1.405
Reference
1000 −1.41
Fcl [N]
p1
500 −1.415
0 −1.42
0 0.02 0.04 0 0.02 0.04
6 0.05
θm [rad]
p2
0
2
0 −0.05
0 0.02 0.04 0 0.02 0.04
300
3
ωm [rad/s]
200
Tm [Nm]
100 0
0 −3
6000 −0.826
Reference
4000
Fcl [N]
p1
−0.828
2000
0 −0.83
0 0.02 0.04 0.06 0 0.02 0.04 0.06
15 0.02
θm [rad]
10 0
p2
5 −0.02
0 −0.04
0 0.02 0.04 0.06 0 0.02 0.04 0.06
300 3
ωm [rad/s]
Tm [Nm]
200 0
100
−3
0
0 0.02 0.04 0.06 0 0.02 0.04 0.06
Time [s] Time [s]
where φe (ωm , θm,f , ωm,f ) denotes the motor position where control switching
should occur, as illustrated in Figure 5.3. Additionally, the instantaneous
106 Chapter 5. High performance clamp force tracking control
300
200
100
ωm [rad/s]
−100
θm,f = 0 10 20 30 40 50
−200
−300
−10 0 10 20 30 40 50 60
φe [rad]
Figure 5.3: Time-optimal switching curves for θm,f = {0, 10, · · · , 50} rad. Solid
lines represent cases for ωm,f = 0 rad/s, and dash-dotted lines represent cases
for ωm,f = 200 rad/s. The velocity constraints are indicated by dash lines.
tracking errors for motor position and velocity are respectively defined by:
e1 = θm − θm,r , (5.17a)
e2 = ωm − ωm,r . (5.17b)
Note that the switching curves can be moved to a common origin as depicted
in Figure 5.4. Using (5.16) and (5.17), the time-optimal feedback control law
can be stated as:
Tm,cv if |ωm | = ωm max and se e2 < 0,
Tm,to = −Tm max sgn (e2 ) if sgn (se ) = 0, (5.18)
sgn (s ) otherwise,
−T
m max e
where Tm,cv represents the motor torque (5.15) that maintains constant velocity
at the velocity limit, given by
600
θm,f = 0, 10, …, 50
400
200
e2 [rad/s]
−200
−400
−600
−20 −15 −10 −5 0 5 10 15 20 25
e1 [rad]
Figure 5.4: Overlaid switching curve plotted in the error state plane: θm,f =
{0, 10, · · · , 50} rad for ramp tracking with ωm,f = 0 (solid) and ωm,f =
200 rad/s (dash-dot). The thick solid curve corresponds to (θmf , ωm,f ) = (0, 0).
6000
Reference
4000
Fcl [N]
2000
0
0 0.02 0.04 0.06 0.08 0.1
400
ωm [rad/s]
200
−200
0 0.02 0.04 0.06 0.08 0.1
5
Tm [Nm]
Control Chatter
−5
0 0.02 0.04 0.06 0.08 0.1
Time [s]
350
300
250
200
ωm [rad/s]
150
100
Control Chatter
50
−50
0 2 4 6 8 10 12 14
θm [rad]
Figure 5.5: Responses for 1 kN (thin lines) and 5 kN (think lines) steps ob-
tained using the time-optimal feedback control law.
5.2. Robust near-time-optimal tracking control 109
developed. The key features of this design are simple implementation and
rapid tracking performance.
In this section, the perturbed EMB model is first presented, followed by
the NTOC design considering tracking and state constraints. Stability anal-
ysis of the closed-loop error system under modelling and measurement errors
is then executed. In order to aid in practical implementation, a gain tuning
procedure is proposed. This procedure is intuitive to implement and allows a
balanced choice between tracking speed and energy usage to be made. Then,
the controller implementation routines are summarised. Lastly, the controller
is implemented on the production-ready prototype EMB, and tested over dif-
ferent braking scenarios to assess the performance and robustness relative to
the benchmark controllers.
Assumption 5.1 (Sufficient control authority in the plant). For a given max-
imum motor torque Tm max , the perturbation δ is small such that there exist a
5.2. Robust near-time-optimal tracking control 111
x(t0 ) = x0 ,
e(t) := x(t) − xr (t),
|e(τ )| ≤ β, τ ≥ tf .
112 Chapter 5. High performance clamp force tracking control
(SC-1) φa (0) = 0,
(SC-2) e2 φa (e2 ) < 0, ∀e2 6= 0,
(SC-3) ξ(e2 ) < 0, where ξ(e2 ) = ∂φ∂e
a (e2 )
2
,
(SC-4) For a given ab , there exist ar , e2δ ≥ 0 such that |e2 | + (ar + ab )ξ(e2 ) > 0
for all |e2 | ≥ e2δ .
5.2. Robust near-time-optimal tracking control 113
Remark 5.3. Figure 5.4 shows the exact switching curves approximated using
a single switching curve that corresponds to xf = (0, 0). The approximated
switching curve crosses the origin, therefore (SC-1) holds. Furthermore, (SC-
2) and (SC-3) hold because the curve passes through the second and fourth
quadrant in the phase plane, and has negative gradient everywhere. Condition
(SC-4) can be used to determine the limiting characteristics of the reference.
The design variable, ab represents the motor acceleration when the velocity is
in vicinity of its constraint. For a particular set of ab and ar , vr has to be
chosen such that ωm max − δ − vr ≥ e2δ . This is to ensure the convergence of
e1 − φa (e2 ) in vicinity of the state boundary.
Having the approximated switching curve defined, the motor position offset
before control switching is given by
sa = e1 − φa (e2 ). (5.23)
e2
II(a): Tm = Tm,cvs
e2 max
e2 max − ∆
IV(b): Tm = −Tm max
e2 (t)
sa = e1 − φa (x2 ) e2sat
−e2sat
II(b): Tm = Tm,cvs
Figure 5.6: Control magnitude in different regions of the error state plane.
The control magnitude over different region in the state plane is shown in
Figure 5.6, and the regions in error state space where control laws (5.24) and
(5.25) operate are listed in Table 5.1.
it is shown that any states in Regions IV(a) and IV(b) (refer to Figure 5.6)
will converge towards e2 = 0, i.e. falling into either Regions III(a) or III(b)
or I. Furthermore, any states in Regions III(a) or III(b) will converge towards
sa = 0, i.e. falling into either Regions II(a), II(b) or I. Then, trajectory
starting in II(a) and II(b) converges towards the switching curve, sa = 0.
Finally, when the error trajectory enters Region I, the asymptotic stability of
a uniform ultimate bound that contains the origin is proved.
Lemma 5.3. Consider the system (5.20) with initial conditions that satisfy
sa e2 > 0 and sa 6= 0. Suppose Assumptions 5.1, 5.2 and 5.3 hold true. Under
the control law (5.26), the trajectories converge to the e1 axis.
Proof. When the states satisfy sa e2 > 0 and sa 6= 0, the control magnitude
is at its maximum limit, that is Tm = −Tm max sgn(sa ). Define a Lyapunov
candidate function V1 = 12 e22 . Its time-derivative is given by
V̇1 = e2 ė2
= e2 (ω̇m − ω̇m,r )
= e2 (−|ω̇m | sgn (sa ) − ω̇m,r )
where use is made of the fact that (5.17b) and there is sufficient control au-
thority to control the sign of ω̇m (from Assumption 5.1). Furthermore, the
bound of the acceleration can be obtained from Assumption 5.1 when the con-
trol magnitude is saturated, that is |ω̇m | ≥ ap + δ. This leads to the following
116 Chapter 5. High performance clamp force tracking control
expressions
where use is made of the fact that e2 sgn (sa ) > 0, |ω̇m,r | ≤ ar (from As-
sumption 5.2) and −e2 ω̇m,r ≤ e2 sgn (sa )ar . Since V̇1 < 0 for all e2 6= 0, the
trajectories converge to e2 = 0, i.e. the e1 axis.
Lemma 5.3 indicates that trajectories with initial condition in Region IV(a)
or IV(b) will move into Regions I, III(a) or III(b).
Lemma 5.4. Consider the system (5.20) with initial conditions that satisfy
sa e2 ≤ 0 and |ωm | < ωm max − ∆. Suppose Assumptions 5.1, 5.2 and 5.3 hold
true, and let
Under the control law (5.26), |sa (t)| will be strictly decreasing, and the set
{|sa | : |σ| ≤ 1} is positively invariant.
V̇2 = sa ṡa
∂φ(e2 )
!
= sa ė1 − · ė2
∂e2
= sa e2 − ξ(e2 )(ω̇m − ω̇m,r )
= sa e2 − sa ξ(e2 ) − |ω̇m | sgn (sa ) − ω̇m,r ,
5.2. Robust near-time-optimal tracking control 117
where use is made of the fact that there is sufficient control authority to control
the sign of ω̇m (from Assumption 5.1). Furthermore,
where use is made of the fact that sa e2 ≤ 0, ξ(e2 ) < 0 (from Assumption 5.3),
|ω̇m | ≥ ap + δ (from Assumption 5.1 with saturated control) and sa ω̇m,r ≥
−|sa |ar (from Assumption 5.2). Therefore, |sa (t)| is strictly decreasing and
the set {sa : |σ| ≤ 1} is positively invariant.
Lemma 5.4 shows that any trajectories within Regions III(a) or III(b) have
strictly decreasing |sa (t)|. However, the state trajectory may reach Region II
(which contains the state constraint) before reaching its invariant set, which
is Region I. The following lemma shows that |sa (t)| is strictly decreasing if the
control law (5.26) is applied when the state is in Region II.
Lemma 5.5. Consider the system (5.20) with initial conditions that satisfies
sa e2 ≤ 0 and |ωm | ∈ [ωm max − ∆, ωm max ]. Suppose Assumptions 5.1, 5.2 and
5.3 hold true. Under the control law (5.26), |sa (t)| will be strictly decreasing.
V̇2 = sa ṡa
= sa e2 − ξ(e2 )(ω̇m − ω̇m,r )
= sa sgn (e2 ) |e2 | + (ω̇m,r − ω̇m )ξ(e2 ) sgn (e2 )
≤ sa sgn (e2 ) |e2 | + (ar + ab )ξ(e2 )
< 0, ∀sa 6= 0
where (SC-4) of Assumption 5.3 is used. Therefore |sa (t)| is strictly decreasing
and reaching the switching curve sa = 0.
Apart from using Lemma 5.5 to show that the |sa (t)| is strictly decreasing in
Region II, the following lemma also shows that the control law (5.26) maintains
the velocity within its constraint in Region II.
118 Chapter 5. High performance clamp force tracking control
Lemma 5.6. Consider the system (5.20) and the control law (5.26), with
initial conditions that satisfy sa e2 ≤ 0 and |ωm | ∈ [ωm max − ∆, ωm max ]. The
region |ωm | ∈ [ωm max − ∆, ωm max ] is uniformly bounded.
Finally, the following lemma shows that the error trajectories in Region I
converge to a uniform ultimate bound that contains the origin.
Lemma 5.7. Consider the perturbed system (5.20), and given that the initial
condition is in the vicinity of the tracking reference, that is sa (0) ∈ {sa :
|σ| ≤ 1}, x ∈ X − {x : |ωm | ∈ [ωm max − ∆, ωm max ], sa ωm < 0}, e2esat 2
≤
N Fcl +T̄f +J ω̇m,r N Fcl +T̄f +J ω̇m,r
1+ Tm max
and e2
e2 sat
≥ −1 + Tm max
. Under the control law
(5.26), the trajectories are uniformly ultimately bounded.
Proof. In vicinity of the error state space e1 -e2 origin, the control law (5.26)
reduces to a linear feedback law,
where
0 1 0 Tm max
A = bks , g(t, e) = bks , b= .
− ssat − e2bsat + sbksats ξ(0) ssat
ζ(e 2 ) + d(t) J
5.2. Robust near-time-optimal tracking control 119
The approximated switching curve is linearised at the origin of the error state,
where φa (e2 ) = ξ(0)e2 + ζ(e2 ). Furthermore, ζ(e2 ) is bounded, such that
ζ(e2 ) ≤ σke2 k2 for all e2 ≤ k2 . Since A is Hurwitz, for any Q = QT > 0, there
exist a unique P = P T > 0 that satisfies the following Lyapunov equation
(Khalil, 2002, Theorem 4.6):
P A + AT P = −Q.
Remark 5.5. Lemma 5.7 shows that the ultimate bound β is proportional to
the upper bound on the disturbance, δ. This result can be interpreted as a
robustness property of the unperturbed system (5.1) with control law (5.26), as
an arbitrarily small δ will result in a small ultimate bound, deviated from the
origin of the error state-space. Although the ultimate bound can be calculated
qualitatively, it is usually found to be too conservative to be used for tuning
of a practical controller (Khalil, 2002).
120 Chapter 5. High performance clamp force tracking control
Collecting the results from Lemma 5.3 to Lemma 5.7, the convergence
analysis can be summarised as the following theorem.
Theorem 5.8. Consider the perturbed EMB system (5.20). Suppose Assump-
tions 5.1, 5.2 and 5.3 hold true, the near-time-optimal tracking control law
(5.26) asymptotically stabilises the tracking error to a uniform ultimate bound
that contains the origin.
Theorem 5.8 indicates that the error trajectory is bounded if the given
assumptions are satisfied. By verifying these assumptions, the applicability of
the NTOC on a given plant can be determined, and the design limits of the
tracking reference can be decided. Furthermore, the quantitative value of the
ultimate bound (5.29) is found to be too conservative to be used as a control
design criteria. Therefore, another design criteria that considers tracking speed
and control effort is proposed in the next subsection.
During EMB clamping operation, the motor position reference θm,r (t) is avail-
able online from the driver input, but the velocity ωm,r (t) and acceleration
ω̇m,r (t) commands are not explicitly available. While ωm,r (t) and ω̇m,r (t) can
be numerically approximated utilising the backward Euler differentiation, this
approach introduces noise and phase offset, which are undesirable in applica-
tion. Based on these practical considerations, a special case of the NTOC is
considered, whereby (5.24) is adapted for step reference tracking, given by:
!
sa ωm N Fcl + T̄f
Tm,step = −Tm max sat ks sat + sat − , (5.30)
ssat ωm sat Tm max
Gain Tuning
The tuning parameters, ks , ssat and ωm sat explicitly represent the size of the
boundary layer (refer to Figure 5.61 ). However, it is unclear how these param-
eters affect the tracking speed and control effort of the controller. Motivated
by the need to address the aforementioned problems and to aid in design, a
systematic gain tuning procedure is proposed. This procedure provides the de-
signer an intuitive tool to choose a set of controller gains that allows a balanced
choice between these conflicting objectives to be made.
Recall that the aims of the NTOC is to provide a chatter free controller with
short rise-time and high-bandwidth. The short rise-time criteria is achieved
by utilising an approximated time-optimal switching curve to determine the
instance to decelerate the motor. To address the high bandwidth criteria, a
local analysis of the closed-loop system is carried out. The local analysis is
sufficient when the trajectory is close to the reference, which is a typical sce-
nario in practise when the reference changes slowly compared to the dynamics
of the system.
Consider the case when the state is in vicinity of zero tracking error and
the control magnitude is unsaturated in the absence of plant disturbance, the
closed-loop dynamics can be calculated using (5.1) and (5.31), governed by
θ̇
m =
0 1 θ
m +
0
θm,r + ζ(ω m , (5.32)
)
ω̇m − sbksats − x2bsat + sbksats ξωm (0) ωm bks
ssat
where b = Tm max
J
and the linearised switching curve is employed, given by
0
m
−2
−4
−300 −200 −100 0 100 200 300
ζ(ωm) = φa(ωm) − ξω (0)ωm [rad]
2
m
−2
−4
−300 −200 −100 0 100 200 300
ωm [rad/s]
Figure 5.7: Comparison of switching curve with its first order linear approxi-
mation.
θ̇
m =
0 1 θm 0 0
+ 1 Tm + , (5.35)
ω̇m 0 0 ωm J
bξω m (ω m )
The relationship between the new gains kp , kd and the original tuning param-
2
Can be verified to be small.
5.2. Robust near-time-optimal tracking control 123
eters are
bks
ssat = , (5.37a)
kp
!−1
kd ks
ωm sat = + ξ(0) . (5.37b)
b ssat
Since the tuning procedure now translates to the selection of a weighting con-
trasting tracking error and control effort, it is more transparent than direct
alteration of the boundary layer size.
Note that techniques such as loop-shaping can be applied for the tuning
of ρ and γ. Consider the open-loop transfer-matrix of (5.35) from the plant’s
input to the controller’s output, larger ρ results in the decrease of the gain
magnitude and slower response. Additionally, larger γ leads to a larger phase
margin and smaller overshoot in the step response.
Figure 5.8 illustrates the trend of ssat and ωm sat with respect to ρ and γ.
Increased penalty on control effort, ρ leads to a larger ssat and also a larger
boundary layer. Furthermore, the weight on motor velocity tracking error γ
has no apparent effect on ssat as all four γ = {0.001, 0.02, 0.03, 0.05} produce
similar ssat that visually stack on top of each other. A decreased γ results in
the stretching of boundary layer vertically. However, for a small γ = 0.001,
the associated value for ωm sat is negative and does not satisfy the condition
ωm sat > 0 needed for the stability of the controller, therefore this choice of γ
is invalid.
The sizes of unsaturated control region produced from different values of
ρ and γ are shown in Figure 5.9. Due to the negative ωm sat associated with
γ = 0.001, the unsaturated control regions are distinctive to the intended de-
sign shown in Figure 5.6, and the resulted control trajectories may be unstable.
124 Chapter 5. High performance clamp force tracking control
6
ssat [rad]
2
γ=0.001 γ=0.02 γ=0.03 γ=0.05
0
1 2 3 4 5 6 7 8 9 10
−8
x 10
300
200
x2sat [rad/s]
100
−100
1 2 3 4 5 6 7 8 9 10
ρ x 10
−8
Figure 5.8: Controller parameters ssat and ωm sat are plotted against ρ and γ
with ks = 2.
Comparing the three figures where γ = 0.01, larger ρ yields a larger unsat-
urated region. Furthermore, comparing the figures with ρ = 1 × 10−8 and
γ = {0.01, 0.05}, larger value of γ reduces the unsaturated region in vicinity
of the origin.
Implementation Procedure
0 0 0
0 0 0
the curve cuts through the second and fourth quadrants, and has negative
gradient everywhere.
Step 3: Determine the design limits of reference. The class of refer-
ence with with bounded higher derivatives is considered, where vr =
270 rad/s and ar = 4300 rad/s (Assumption 5.2). For the unper-
turbed case, where δ = 0, the bottom plot of Figure 5.10 shows that
|e2 | + ar ξ(e2 ) > 0 for all |e2 | ∈ [e2δ , 650], e2δ = 29 rad/s. Hence, (SC-4)
of Assumption 5.3 is satisfied.
Step 4: Controller synthesis. The EMB clamp force tracking controller is
implemented as shown in Figure 5.11. The clamp force reference and
measurement are first converted to the motor position reference and mea-
surement respectively. Then, the motor position and velocity references,
together with their measurements are fed into the control law (5.31).
Step 5: Gain tuning. The controller gains ks , ssat , ωm sat are tuned using the
procedure proposed in this subsection.
Simulation results
Initially, results obtained via simulation are used to compare the different
control approaches. The investigation begins with the selection of NTOC
gains, followed by the comparison of different controllers on tracking response.
5.2. Robust near-time-optimal tracking control 127
20
5th order polynomial fit
Switching curve
10
φa(e2)
−10
−20
0
ξ(e2)
−0.01
−|e2|/ar
ξ(e2), −|e2|/ar
−0.02
−0.03
−0.04
−0.05
−800 −600 −400 −200 0 200 400 600 800
e2 [rad/s]
Figure 5.10: (Top) Approximation of the switching curve using a 5th order
polynomial. (Bottom) For ar = 4300 rad/s2 , the first order differentiation of
the switching curve, ξ(e2 ) is larger than − |ea2r | for all 29 ≤ |e2 | ≤ 650 rad/s.
θm,r
EMB
Fcl,r Convert force to θm Simplified NTOC Tm Control-
Fcl
motor position Tm,ntocs Oriented
Model
Fcl ωm
Gain selection. As proposed in Section 5.2.4, the NTOC gain tuning can
be translated to the selection of weights contrasting control effort and tracking
error, given by ρ and γ. To make an informed selection for the gains, the
control performances using a range of weights are compared.
Figure 5.13(a) compares 2 kN step responses with different weighting on
the control effort, ρ. A smaller value of ρ leads to a faster response, but it is
also more likely to induce control chatter as demonstrated in the case where
ρ = 0.1 × 10−8 . This is because a small value of ρ produces a small ssat ,
therefore leading to a narrower boundary layer. The NTOC with an infinitely
small boundary layer has an abrupt change in control magnitude across the
switching curve, which is also observed in the time-optimal control and can
5.2. Robust near-time-optimal tracking control 129
γ = 0.05
3000
2000
Fcl [N]
1000 ρ=0.1×10−8
ρ=2.5×10−8
0
−8
0 0.02 0.04 0.06 0.08 ρ=5×10
0.1 0.12
−8
400 ρ=10×10
ωm [rad/s]
200
−200
0 0.02 0.04 0.06 0.08 0.1 0.12
5
Tm [Nm]
−5
0 0.02 0.04 0.06 0.08 0.1 0.12
Time [s]
ρ = 5 × 10−8
4000
γ = 0.001 γ = 0.02 γ = 0.03 γ = 0.05
Fcl [N]
2000
0
0 0.05 0.1 0.15 0.2
400
ωm [rad/s]
200
−200
0 0.05 0.1 0.15 0.2
5
Tm [Nm]
−5
0 0.05 0.1 0.15 0.2
Time [s]
0.08
Time−optimal control γ = 0.2 γ = 0.3
0.07
Rise Time [s]
0.06
0.05
0.04
1 2 3 4 5 6 7 8 9 10
−8
x 10
25
Energy Usage [J]
20
15
10
1 2 3 4 5 6 7 8 9 10
ρ −8
x 10
Figure 5.14: Rise time and energy usage for a 2 kN step using the time-optimal
control and NTOC with γ = {0.02, 0.03} and ρ ∈ [1, 10] × 10−8 .
However, for high frequency (> 10 Hz), the gain magnitude varies significantly
with clamp force levels. In particular, for a small clamp force level at 1 kN,
the gain magnitude deviates from 0 dB and start decreasing at approximately
25 Hz. For a larger clamp force level at 30 kN, the decrease starts at ap-
proximately 10 Hz. The decrease in gain magnitude starts earlier with higher
clamp force because higher clamp force escalates both the load torque and fric-
tion torque, therefore reduces the achievable acceleration needed to maintain
a small tracking error.
Compared to the general NTOC, the simplified NTOC has smaller gain
magnitudes and larger phase-shifts, as depicted in Figure 5.15(b). This is due
to the removal of motor velocity reference in the control law.
Figure 5.16 shows the gain and phase-shift of the clamp force output rel-
ative to the reference with the industry-tuned baseline PI controller. Unlike
the general NTOC, the gain magnitudes with the PI are inconsistent for for
132 Chapter 5. High performance clamp force tracking control
0
Magnitude (dB)
−5
−10 1 kN
5 kN
10 kN
−15 20 kN
30 kN
−20
0
−45
Phase (deg)
−90
−135
−180
−225
0 Frequency (Hz) 1
10 10
−5
Magnitude (dB)
−10
−15 1 kN
5 kN
10 kN
−20 20 kN
30 kN
−25
0
−45
Phase (deg)
−90
−135
−180
−225
0 1
10 10
Frequency (Hz)
Figure 5.15: Gain and phase-shift of the clamp force output relative to the
reference with NTOCs.
5.2. Robust near-time-optimal tracking control 133
−20
1 kN
5 kN
−30 10 kN
20 kN
30 kN
−40
0
−45
Phase (deg)
−90
−135
−180
0 1
10 10
Frequency (Hz)
Figure 5.16: Gain and phase-shift of the clamp force output relative to the
reference with baseline PI controller.
all force levels. In particular, the gain magnitudes for smaller clamp force
levels are typically smaller or equal compared to the higher clamp force lev-
els. This reflects a worse tracking performance at small clamp force levels,
which is resulted from not fully utilising the available motor capacity at small
clamp force levels. Additionally, the baseline PI controller produces responses
with larger gain magnitudes and phase-shifts compared to NTOC, therefore
indicating that the NTOC has a better tracking performance.
For the purpose of the robustness study, α ∈ [0.5, 1.5] is randomly sampled
from a uniform distribution. Similarly, the friction parameters are also ran-
134 Chapter 5. High performance clamp force tracking control
Experimental results
The experimental results presented here were performed using the static test
rig of the prototype EMB described in Chapter 3. The simplified NTOC (5.31)
was implemented to the onboard 16-bit microcontroller, where the clamp force
control loop operates at 250 Hz. The NTOC gains ks = 2, ssat = 4 and
e2 sat = 50 were used
The performance of the EMB controllers are assessed experimentally by
the rise time and how well they tracks a reference, which are all important
for demanding situations such as panic braking and during ABS operations.
These criteria will be assessed by the execution of tests using standardised
waveforms such as step, sinusoidal and triangle. Furthermore, performance of
the controllers will also be discussed in terms of energy usage.
Rise time. Step tracking performance of the controllers are first assessed
using a large step, and then followed by a smaller step. The large step ma-
noeuvre corresponds to a full brake application in emergency braking while the
small step demand corresponds to a light brake application. A faster step re-
sponse improves reaction time in emergency braking and may reduce stopping
distance. Figure 5.18(a) illustrates the large step response of 20 kN, where
both the proposed NTOC and the baseline PI controller have the same rising
slope and similar rise time, indicating that the PI is well-tuned for a large
manoeuver. However, the PI is slower at bringing the response to 0 kN, as the
input switches from negative to positive too early. The NTOC responses are
close to the ideal case apart from some overshoot during the falling phase, due
5.2. Robust near-time-optimal tracking control 135
4000
Fcl [N] Perturbed Model Nominal Model
2000
0
0 0.05 0.1 0.15 0.2 0.25 0.3
400
ωm [rad/s]
200
−200
0 0.05 0.1 0.15 0.2 0.25 0.3
100
Tm [Nm]
−100
0 0.05 0.1 0.15 0.2 0.25 0.3
Time [s]
4
x 10
4
Perturbed Model Nominal Model
Fcl [N]
0
0 0.05 0.1 0.15 0.2 0.25 0.3
400
ωm [rad/s]
200
−200
0 0.05 0.1 0.15 0.2 0.25 0.3
100
Tm [Nm]
−100
0 0.05 0.1 0.15 0.2 0.25 0.3
Time [s]
Figure 5.17: Responses of the nominal and the perturbed plant models.
136 Chapter 5. High performance clamp force tracking control
Although the proposed NTOC performed satisfactorily for a full brake ap-
ply, to ensure the consistency of step tracking performance within the range
of operation, a smaller step manoeuvre of 2 kN is performed, and is shown in
Figure 5.18(b). The rising and falling responses of the NTOC are faster com-
pared to the PI, suggesting that the fixed set of PI gains are too conservative
for a small step reference. As shown in the measured velocity and motor cur-
rent, the NTOC fully utilises the available motor torque to track the setpoint
reference.
To assess how close the NTOC performs relative to the limitation of the
plant, the 0%–95% rise time of the NTOC is compared to the time-optimal
control, as depicted in Figure 5.20. For smaller steps, notably for steps below
2 kN, the rise time of the NTOC is significantly longer compared to the time-
optimal control. This is due to the introduction of the boundary layer in the
NTOC to avoid control chatter. However, the rise times of the NTOC for
larger steps are within 5% of the time-optimal control, indicating that the
NTOC is able to fully utilise the available motor capacity to achieve a rapid
step tracking. Note that the rise time of the baseline PI is always longer
compared to the NTOC over the whole clamp force operating range.
5.2. Robust near-time-optimal tracking control 137
4
x 10
4
Ref. PI (exp.) NTOC (exp.) NTOC (sim.)
Fcl [N]
2
0
0 0.5 1 1.5
300
ωm [rad/s]
−300
0 0.5 1 1.5
3
Tm [Nm]
−3
0 0.5 1 1.5
Time [s]
4000
Ref. PI (exp.) NTOC (exp.) NTOC (sim.)
Fcl [N]
2000
0
0 0.5 1 1.5
300
ωm [rad/s]
−300
0 0.5 1 1.5
3
Tm [Nm]
−3
0 0.5 1 1.5
Time [s]
400
Velocity Constraint
300
200
100
ωm [rad/s]
−100
−200
2 kN: Basline PI Switching Curve
−300 2 kN: NTOC
20 kN: Basline PI Velocity Constraint
20 kN: NTOC
−400
−35 −30 −25 −20 −15 −10 −5 0 5 10 15
e1 [rad]
0.4
Time−optimal control (simulation)
NTOC (simulation)
0.35 NTOC (experiment)
Baseline PI (experiment)
0.3
0.25
Rise time [s]
0.2
0.15
0.1
0.05
0
0 0.5 1 1.5 2 2.5 3
Fcl,r [N] 4
x 10
Figure 5.20: Rise time for step responses initiated from 0.4 kN.
5.2. Robust near-time-optimal tracking control 139
4
x 10
Fcl [N] 2
4000
Tracking error [N]
2000
−2000
−4000
0 0.5 1 1.5 2 2.5 3
Time [s]
where e, Fcl,r(mean) and T respectively represent the clamp force tracking error,
the mean and the period of the clamp force reference trajectories.
Figure 5.21 compares the tracking performance of the baseline PI controller
and the NTOC, where a triangular profile from 0 to 20 kN with 1 second period
was commanded. For the 40 kN/s ramp found in the triangular profile, the
NTOC demonstrates an approximately threefold improvement in %nIAE as
listed in Table 5.2.
Figure 5.22 compares the sinusoidal reference tracking performance of both
controllers. It is shown that the simplified NTOC has a larger bandwidth, and
visually appears to track the reference significantly better than the baseline
PI controller. There is a corresponding 65% reduction in %nIAE, as listed in
Table 5.2.
140 Chapter 5. High performance clamp force tracking control
3000
Ref. PI NTOC (5.29) NTOC (5.24)
2000
Fcl [N]
1000
0
0 0.5 1 1.5 2 2.5 3
Tracking error [N]
1000
−1000
0 0.5 1 1.5 2 2.5 3
3
Tm [Nm]
−3
0 0.5 1 1.5 2 2.5 3
Time [s]
Additionally, the tracking response of the general NTOC (5.26) is also plot-
ted in Figure 5.22, where it shows great reduction in tracking error compared to
the PI and the simplified NTOC (5.31), and with the %nIAE of 3.02%. How-
ever, due to the employment of numerical differentiation for state reference,
the torque command is very noisy, and this may be an important consideration
for its implementation. On the other hand, the noise level found in the torque
command from the simplified NTOC is within an acceptable level.
Energy Usage. The energy usage of the baseline PI and NTOC is shown in
Figure 5.23. In order to achieve a shorter rise time, NTOC uses more energy to
achieve higher acceleration and deceleration compared to the PI. Nevertheless,
it is noted that the gain tuning proposed in Section 5.2.4 is flexible enough to
5.3. Explicit model predictive control 141
70
Baseline PI
NTOC
60
50
Energy Usage [J]
40
30
20
10
0
2 5 10 15 20 25 30
Fcl,r [kN]
Figure 5.23: Experimentally obtained energy usage for step responses initiated
from 0 N clamp force.
Tm (t) ∈ U,
x(t) ∈ X ,
x(t0 ) = x0 .
Problem 5.3 (Model predictive clamp force control). Given an initial state,
x(k) = x0 , a prediction horizon Hp , a control horizon Hu ≤ Hp , a clamp
force reference up to the prediction horizon, {Fcl,r (k + i), i = 1, . . . , Hp } and
a sampling period ts , find the optimal change in torque trajectory up to the
control horizon, ∆U ∗ := {∆Tm∗ (k + i), i = 1, . . . , Hu } such that the weighted
cost of tracking error and control deviation is minimised, subjected to the state
5.3. Explicit model predictive control 143
i=1 i=1
subject to
xd (k + i + 1) = xd (k + i) + ts f xd (k + i), Tm (k + i), θ ,
(5.42a)
yd (k) = h xd (k) , i = 0, . . . , Hp
Tm (k + i) = Tm (k + i − 1) + ∆Tm (k + i), i = 1, . . . , Hp (5.42b)
∆Tm (k + i) = 0, i = Hu + 1, . . . , Hp
Tm (k + i) ∈ U, i = 0, . . . , Hp
xd (k + i) ∈ X , i = 0, . . . , Hp
xd (k) = x0 .
The discretised model (5.42a) has been derived in (3.24)–(3.25), where the
function f (·) and h(·) are given in (3.23). Furthermore, control deviation is
considered here as a means of incorporating integral action within the MPC
framework to ensure offset-free tracking in the face of model uncertainties and
unknown constant disturbances (Maciejowski, 2002). The control prediction
Hp typically needs to be large to ensure stability, although extending Hp has
only minimal effect of the complexity of solving the MPC problem. For good
tracking performance, Hp is chosen large enough to include the transients of the
dominant plant dynamics. On the other hand, extending the control horizon
Hu will increase the complexity significantly. The choice of Hu is a tradeoff
between tracking precision and computation effort. Note the predicted control
input (5.42b) remains unchanged beyond the control horizon.
Figure 5.24: Motor torque deviation map with fixed reference at 13 kN.
where the change in control action at time step k is a function of the reference
Fcl,r (k), measured outputs Fcl (k), ωm (k) and the previous control input Tm (k −
1).
The optimum torque deviation ∆Tm (k) is depicted in Figure 5.24 as a
function of clamp force Fcl (k), motor speed ωm (k) and motor torque Tm (k −1),
where the clamp force reference Fcl,r (k) is fixed at 13 kN. The control map
∆Tm,mpc (·) is discretised with 31 sections of clamp force reference Fcl,r from
0 kN to 30 kN, 34 sections of clamp force Fcl from 0 kN to 33 kN, 23 sections
of motor speed from -330 rad/s to 330 rad/s, and 26 sections of motor torque
Tm from -3 Nm to 3 Nm. Online implementation of the control law (5.43) is
achieved using a lookup table, where the binary search method is utilised to
locate the optimal deviation in control action ∆Tm (k).
5.3. Explicit model predictive control 145
0.1
Hu = 2
Rise Time [s]
Hu = 3
0.08
Time−optimal control
0.06
0.04
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25
80
Maximum Overshoot [%]
60
40
20
−20
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25
Hp
0.25
Hu = 1
Rise Time [s]
Hu = 2
0.2
Hu = 3
Time−optimal control
0.15
0.1
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25
15
Maximum Overshoot [%]
10
−5
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25
Hp
0.09
Time−optimal control
0.08 WQ,d = 1
Rise Time [s]
0.06
0.05
0.04
−2 −1 0 1 2 3 4
10 10 10 10 10 10 10
25
Energy Usage [J]
20
15
10
5
−2 −1 0 1 2 3 4
10 10 10 10 10 10 10
WR,d / WQ,d
Figure 5.26: Rise time and energy usage for a 2 kN step using the time-optimal
control and MPC with WQ,d = {1, 100} and WR,d /WQ,d ∈ [10−2 , 102 ].
tively. For both manoeuvres, the time-optimal control is the quickest to reach
the setpoints, but it experiences control chatter when the setpoints are reached.
Both NTOC and MPC are free from chatter. However, the MPC has larger
control deviations compared to the NTOC that can quickly bring the clamp
force trajectory to the setpoint. Although a properly tuned MPC potentially
achieves shorter rise time compared to a NTOC tuned for robustness, the
improvements are visually insignificant in these graphs.
24
MPC WQ,d = 1
22 MPC WQ,d = 100
NTOC γ = 0.2
20 NTOC γ = 0.3
Time−optimal control
18
Energy Usage [J]
16
14
12
10
6
Increasing WR,d / WQ,d (for MPC) or ρ (for NTOC).
4
0.045 0.05 0.055 0.06 0.065 0.07 0.075 0.08 0.085
Rise Time [s]
Figure 5.27: Comparison of rise time and energy usage of the time-optimal
control, NTOC and MPC.
the optimal control deviation from a lookup table, which was prepared offline
for a bounded operating region. The simulation scenario was 0.2 seconds of
13 kN step clamp force response. It was run on a 2.8 GHz Pentium 4 machine
where Simulink were used with a variable-step solver ode45 and the maxi-
mum step of 0.0002 seconds. It is observed that the online MPC requires a
very long simulation time, and is prohibitive for practical implementation on
the EMB microcontroller.
On the other hand, the explicit MPC requires significantly shorter com-
putation time, where majority of the computation time is spent on searching
for the corresponding control input on the lookup table. In order to reduce
the interpolation error of the control input, a finer grid for the control law
domain is desired. However, the size of control law lookup table increases as
the domain is discretised finer, whereby larger controller memory is needed.
Although the explicit MPC alleviates the lengthy computational issue by
150 Chapter 5. High performance clamp force tracking control
4000
Time−optimal control NTOC MPC
Fcl [N]
2000
0
0 0.02 0.04 0.06 0.08 0.1 0.12
400
x2 [rad/s]
200
−200
0 0.02 0.04 0.06 0.08 0.1 0.12
5
Tm [Nm]
−5
0 0.02 0.04 0.06 0.08 0.1 0.12
Time [s]
4
x 10
4
Time−optimal control NTOC MPC
Fcl [N]
0
0 0.05 0.1 0.15 0.2
400
x2 [rad/s]
200
−200
0 0.05 0.1 0.15 0.2
5
Tm [Nm]
−5
0 0.05 0.1 0.15 0.2
Time [s]
Table 5.3: Time required for simulating a 0.2 seconds trajectory with a 13 kN
setpoint.
MPC Algorithms Simulation Time [s]
Online MPC 255
Explicit MPC 0.177
Without controller 0.104
performing the optimisation offline and storing the results in a lookup table, it
is found that the storage requirement of the lookup table is prohibitive under
current hardware limits. This suggests the adoption of NTOC with reduced
storage requirements for practical implementation.
5.4 Conclusions
High performance clamp force tracking controllers for the prototype automo-
tive electromechanical brake (EMB) are developed in this chapter. The con-
trollers are designed such that it can track a given reference closely in the
face of nonlinear dynamics and system constraints, where the input and state
constraints are the motor torque saturation and motor torque velocity limits
respectively. A good tracking performance is desirable to meet demanding
requests from higher level controllers, such as the anti-lock braking system
(ABS) and the electronic stability control (ESC). The main contributions of
this chapter are summarised as the following:
EMB
Clamp Force Tm Control-
Fcl
Fcl,r Controller Oriented
Model
Fcl ωm
Figure 5.29: EMB closed-loop system and clamp force controllers developed
in this chapter.
The EMB closed-loop system and the clamp force feedback controllers in-
vestigated in this chapter are illustrated in Figure 5.29. In particular, the
achieved rapid tracking performance and high closed-loop bandwidth of the
NTOC allows additional branching problems to be investigated. In order to
attenuate brake judder, brake torque variation compensators building on the
outer loop of the clamp force controller will be investigated next.
6
155
156 Chapter 6. BTV compensation with wheel speed measurements
Tb Fcl ωm
Note that ωd := θ̇d and θd are the disc rotor angular speed and position re-
spectively.
Note that kp and kd can be calculated using the expressions in (5.37), which
are given by
bks b bks
kp = , kd = + ξω (0),
ssat x2 sat ssat m
where ks , ssat , x2 sat are the NTOC gains, ξωm (·) is the gradient of the switching
curve and b = TmJmax . Furthermore, consider the motor position θm can be
approximated by a fixed θ̄m and the clamp force trajectory is in the vicinity of
Fcl (θ̄m ), then a linear approximation of the clamp force can be obtained as
k̄(θ̄m ) = k1 θ̄m
2
+ k2 θ̄m . (6.3)
by applying a clamp force on both surfaces of a disc rotor via brake pads, and
is governed by
Tb = 2µrd Fcl + Tbtv (θd ), (6.4)
where µ, rd and Tbtv (·) are the coefficient of friction between pads and rotor,
effective disc rotor radius and BTV respectively. Exact values of the coefficient
of friction, µ is not available during online implementation and it is affected
by many environmental factors, such as temperature, speed and humidity.
However, its variation is slow relative to the system dynamics (6.1), thus the
discrepancy between the estimated friction coefficient, µ̂ and the true friction
coefficient, µ can be assumed constant. This leads to the perturbed form of
(6.4), given by
Tb = 2µ̂rd Fcl + Td + Tbtv (θd ), (6.5)
where Td is the discrepancy in brake torque resulted from the estimated error
in friction coefficient.
The BTV in (6.5) can be approximated using the following grey-box model:
n
Tbtv (θd ) = ab,k cos (kθd + φb,k ), (6.6)
X
k=1
where
wk,1 0 kωd
wk = , Sk (ωd ) = . (6.8)
wk,2 −kωd 0
The sum of (6.7b) for all harmonic from 1 to n renders an equivalent grey-box
model of the BTV, which is governed by
k=1 | {z }
n pairs.
where
w̆ = [w1T , . . . , wnT ]T , S̆(ωd ) = diag S1 (ωd ), . . . , Sn (ωd ) .
Note S̆(ωd ) is a square matrix comprises main diagonal blocks of Sk (ωd ) and
off-diagonal blocks of zeros.
Employing the EMB closed-loop system model (6.1), the brake torque ex-
pression (6.5) and the BTV grey-box model (6.9), the dynamics of the brake
corner can be represented in the following form:
ẋ
=
A 0 x Bu
+ Fcl,r (6.10)
ẇ 0 S(ωd ) w 0
where x := [θm , ωm ]T is the system state and w := [Td , w1T , . . . , wnT ]T is the
exogenous system state. The matrices in (6.10) are given by
0 1 0
A= , Bu = kp
,
−kp −kd k̄(θ̄m ) (6.11)
S(ωd ) = diag 0, S1 (ωd ), . . . , Sn (ωd ) ,
where Sk (·) is defined in (6.8). Additionally, the brake torque tracking error
can be characterised by
where Tb,r is the brake torque reference and the matrices are given by
h i
C = 2µ̂rd k̄(θ̄m ) 0 ,
(6.13)
h i
Dew = 1 1 0 ··· 1 0 .
| {z }
n pairs.
160 Chapter 6. BTV compensation with wheel speed measurements
˙
x̄
=
A 0 x̄ Bu
+ ū + Bv v,
0 S(ωd ) w 0
ẇ
Σbc : (6.15)
h i x̄
eb = C Dew + Dv v,
w
where xc is the controller state, and Ac , Bc , Cc are chosen to achieve some spe-
cific control objectives that will be defined shortly. Additionally, to facilitate
brake torque tracking, the following feedforward command is defined:
1
Fcl,f r (Tb,r ) = Tb,r . (6.17)
2µ̂rd
The interconnection of the brake corner model (6.15) and the compensator
(6.16) yields the following closed-loop system:
The control objectives are to internally stabilise the closed-loop system (6.18),
to attenuate the brake torque tracking error, and to minimise a quadratic
performance metric. These objectives are formally stated as the following:
Assumption 6.2. The brake corner model (6.15) has the following properties:
Remark 6.1. It can be verified using parameters from Table 4.3 and NTOC
gains chosen in Section 5.2 that the system (6.15) satisfies conditions (BC-1)
and (BC-2). Moreover, these conditions guarantee the internal stability of the
closed-loop system stated in Problem 6.1.
Furthermore, a fundamental matrix of S(·) is Φ(t0 , t) = diag (1, Φ1 , . . . , Φn ),
where
cos (kθd ) sin (kθ )
d
Φk = , k = 1, . . . , n.
− sin (kθd ) cos (kθd )
Note that kΦ(t0 , t)k = 1, therefore (BC-3) is satisfied.
6.2. Design of a BTV compensator with wheel speed measurements 163
Π̇ + ΠS = AΠ + Bu Γ, (6.20a)
lim (CΠ + Dew ) = 0. (6.20b)
t→∞
Substituting (6.11), (6.13) and (6.21) into the steady-state of (6.20b) yields
i Π · · · Π1,2n+1 h
1,1
h i
2µ̂rd k̄ 0 + 1 1 0 · · · 1 0 = 01×(2n+1) . (6.22)
Π2,1 · · · Π2,2n+1 | {z }
n pairs.
For clarity, the size of the zero matrix is shown, where 01×(2n+1) denotes a zero
matrix with one row and 2n + 1 columns. Expansion of (6.22) gives
1 1
Π1,1 = − , Π1,2k = − , Π1,2k+1 = 0, (6.23)
2µ̂rd k̄ 2µ̂rd k̄
where k = 1, . . . , n. Note the k-sequence will be assumed in the following steps
without restating. Taking the time-derivative of (6.23) leads to
Π̇1,i = 0, i = 1, . . . , 2n + 1. (6.24)
Π̇1,1 = Π2,1 ,
Π̇1,2k − Π1,2k+1 (kωd ) = Π2,2k , (6.26)
Π̇1,2k+1 + Π1,2k (kωd ) = Π2,2k+1 .
To obtain the terms for the second row of Π, (6.23) and (6.24) are substituted
into (6.26) to get
kωd
Π2,1 = 0, Π2,2k = 0, Π2,2k+1 = − . (6.27)
2µ̂rd k̄
Differentiating (6.27) provides
k ω̇d
Π̇2,1 = 0, Π̇2,2k = 0, Π̇2,2k+1 = − . (6.28)
2µ̂rd k̄
The results from (6.23) and (6.27) are collected to obtain Π, which is given by
1 −1 −1 0 −1 0 · · · −1 0
Π(ωd ) = . (6.29)
2µ̂rd k̄ 0 0 −ωd 0 −2ωd · · · 0 −nωd
Additionally, the time derivative of Π is given by
1 0 0 0 0 0 ··· 0 0
Π̇ = . (6.30)
2µ̂rd k̄ 0 0 −ω̇d 0 −2ω̇d · · · 0 −nω̇d
Γ(ωd , ω̇d ) = 1
2µ̂rd kp
[ −kp ωd2 −kp −(ω̇d +kd ωd ) ··· (nωd )2 −kp −n(ω̇d +kd ωd ) ]. (6.32)
Since both ωd and ω̇d are bounded (as stated in Assumption 6.1), the solutions
of Π(ωd ) (given in (6.29)) and Γ(ωd , ω̇d ) (given in (6.32)) are also bounded.
Therefore, a compensator that solves Problem 6.1 exists.
6.2. Design of a BTV compensator with wheel speed measurements 165
x̂˙
=
A 0 x̂ Bu Lx h i x̂
+ Fcl,r + C Dew − eb ,
ŵ˙ 0 S ŵ 0 Lw ŵ
(6.33)
h i x̂
Fcl,r = Kx (Γ − Kx Π) + Fcl,f r ,
ŵ
are both Hurwitz. Note the feedforward gain Γ − Kx Π is utilised for com-
pensating the BTV, which enters the brake corner system at the output, as
depicted in Figure 6.1. In the notation of (6.16), xc = [x̂T , ŵT ]T and
A + Lx C + Bu Kx Lx Dew + Bu Γ(ωd , ω̇d ) − Kx Π(ωd )
Ac (ωd , ω̇d ) = ,
Lw C S(ωd ) + Lw Dew
Lx h i
Bc = − , Cc (ωd , ω̇d ) = Kx (Γ(ωd , ω̇d ) − Kx Π(ωd )) .
Lw
166 Chapter 6. BTV compensation with wheel speed measurements
Wheel speed and acceleration estimation. Note both wheel speed and
acceleration are required for implementing the compensator (6.33), however
only the speed measurements are available. Therefore, the acceleration is
estimated from the measured speed, where the following estimator is used:
ω̂˙
d =
0 1 ω̂d
+ Lω (ω̂d − ωd ), (6.35)
˙α̂d 0 0 α̂d
where ω̂d and α̂d are the estimated speed and acceleration respectively. The
estimator gain Lω ∈ R2 is chosen such that ([ 00 10 ] + Lw [ 1 0 ]) is Hurwitz stable.
where co (·) denotes the convex hull. The performance metric (6.36) is devised
based on the separation principle. The first part of the metric, Jbtvc,s1 is used for
optimising the controller gain, Kx with respect to tracking speed and control
effort, whereas the second part of the metric, Jbtvc,s2 is employed to find an
optimal observer gain L that minimises the supremum of the state estimation
error’s expected value over the wheel speed operating range.
Note the system state x̄ is considered, but the exogenous system state w
is excluded from the controller gain Kx selection. This is because the exoge-
nous system ẇ = Sw is decoupled from the input Fcl,r , and hence the pair
([ C Dew ], [ A0 S0 ]) is uncontrollable. Therefore, the controller design cannot be
arranged in the traditional linear quadratic gaussian (LQG) setting, and is
considered using two decoupled objectives, Jbtvc,s1 and Jbtvc,s2 .
The solution for minimising Jbtvc,s1 is provided by the linear quadratic reg-
ulation (LQR) theory, where the following algebraic Ricatti equation is solved
(Dorato et al., 1995):
Kx = R−1 B T Pare .
The contradicting objectives of tracking speed and energy usage can be tuned
using the weight matrices, Q and R.
Furthermore, the solution for minimising Jbtvc,s2 is translated into the fol-
lowing LMI (de Souza and Trofino, 2000):
min tr (N ),
N,P,W
N BwT P + DwT W T
subject to ≥ 0, P > 0,
P Bw + W D w P
ATaug,i P + P Aaug,i + Caug
T
W T + W Caug + LT L < 0, i = 1, 2
(6.38)
where
A 0 A 0 h i
Aaug,1 = , Aaug,2 = , Caug = C Dew .
0 S(ωd ) 0 S(ωd )
45
40
35
Brake Torque Variation [Nm]
30
25
20
15
10
−5
0 1 2 3 4 5 6 6.283
Disc Rotor Angular Position [rad]
140
Tb [Nm]
120
100
80
0 0.5 1 1.5 2 2.5 3 3.5 4
1000
800
600
Fcl [N]
400
200
Uncomp. 1st 2nd 3rd
0
0 0.5 1 1.5 2 2.5 3 3.5 4
Time [s]
Figure 6.3: Simulated responses for BTV compensators with different orders
of harmonic.
insignificant from a practical perspective and does not justify the increased
complexity in implementation. Therefore, compensator that includes up to
the second-order harmonic is anticipated to perform well in practise.
Additionally, the average and peak power requirements of the compen-
sators in steady-state over five wheel revolutions are compared in Table 6.1.
Unsurprisingly, compensators of higher order require more power in order to
achieve a better BTV reduction due to increased motor activity. The peak
power required by the third-order compensator is 50.7 W, which is well within
the 1 kW power rating of the EMB motor.
6.3. Results and discussions 171
140
Tb [Nm]
120
100
80
0 0.5 1 1.5 2 2.5 3 3.5 4
1000
800
600
Fcl [N]
400
200
Q/R = 1 Q/R = 103 Q/R = 106
0
0 0.5 1 1.5 2 2.5 3 3.5 4
Time [s]
Figure 6.4: Simulated brake torque and clamp force trajectories for different
Q/R ratios.
Gain tuning. In order to study the effects of tuning parameters, they are
perturbed one at a time from the nominal set of values, starting from the Q/R
ratio, then the measurement noise matrix Dw , followed by the process noise
matrix Bw . Brake torque and clamp force trajectories for different Q/R ratios
for a second-order compensator are shown in Figure 6.4. A larger Q/R ratio
yields a larger norm on the controller gain Kx . As a result, the feedforward gain
(Γ − Kx Π) decreases and the BTV compensation performance is decreased.
Compensation performance for different magnitudes of measurement noise
Dw are compared in Figure 6.5. Larger measurement noise indicates that the
measurements are unreliable and thus leads to a conservative estimator gain
and a slower compensation response as expected.
To study the effect of process noise on compensation performance, the nom-
inal value of Bw is perturbed in the form of Bw = [0, 1, 1, 0, ν, 0, ν]T , where ν is
the process noise standard deviation for the sinusoidal BTV model. Compen-
sation performance for different magnitudes of ν are compared in Figure 6.6.
A larger observer gain is obtained for a larger process noise, leading to a faster
172 Chapter 6. BTV compensation with wheel speed measurements
140
Tb [Nm]
120
100
80
0 0.5 1 1.5 2 2.5 3 3.5 4
1000
800
600
Fcl [N]
400
200 Dw = 1 Dw = 2 Dw = 10
0
0 0.5 1 1.5 2 2.5 3 3.5 4
Time [s]
Figure 6.5: Simulated brake torque and clamp force trajectories for different
Dw magnitudes.
180
160
140
Tb [Nm]
120
100
80
60
0 0.5 1 1.5 2 2.5 3 3.5 4
1500
1000
Fcl [N]
500
Figure 6.6: Simulated brake torque and clamp force trajectories for different
ν magnitudes, where the process noise is defined as Bw = [0, 1, 1, 0, ν, 0, ν]T .
ts = 0.0001 s, ωd = 3 Hz ts = 0.0001 s, ωd = 10 Hz
Tb [Nm]
120 120 Envelope
100 100
80 80
0 2 4 6 0 0.5 1 1.5 2
ts = 0.001 s, ωd = 3 Hz
ts = 0.001 s, ωd = 10 Hz
140
140
Tb [Nm]
Tb [Nm]
120 120
100 100
80 80
0 2 4 6 0 0.5 1 1.5 2
Time [s] Time [s]
Fixed wheel speed. The first experimental test case corresponds to brak-
ing under fixed wheel speed, which commonly occurs when light braking is
applied to a vehicle descending a ramp. Figure 6.9 shows the BTV compen-
sation for different wheel speeds. Due to the reduction of friction coefficient
over increasing wheel speed, a larger clamp force reference was given (before
compensation) for faster wheel speed. As the compensator was started, the
6.3. Results and discussions 175
ts = 0.0001 s, ωd = 3 Hz ts = 0.0001 s, ωd = 10 Hz
Tb [Nm]
120 Euler 120
100 100
80 80
0 2 4 6 0 0.5 1 1.5 2
ts = 0.001 s, ωd = 3 Hz ts = 0.001 s, ωd = 10 Hz
140 140
Tb [Nm]
100 100
80 80
0 2 4 6 0 0.5 1 1.5 2
amplitude of the brake torque responses were attenuated, and the averages
of the responses approached the setpoint at 100 Nm. In Figure 6.9(a), some
fluctuations are seen around 28 seconds as well as that around 6.5 seconds
in Figure 6.9(b). These fluctuations were caused by dropped packets in the
CAN bus. Note that the BTV algorithm was implemented and tested in Vector
CANape, which runs under Windows XP. Since Windows XP is not a real-time
operating system, packet drop may occur during testing.
Due to the integration action included in the compensator, the mean of the
responses were corrected and moved towards the setpoints when the compen-
sator was started. The normalised RMSEs of the compensated responses are
listed in Table 6.2, indicating that the BTV was attenuated for different wheel
speeds. However, the attenuation performance decreases as the wheel speed is
increased.
To investigate the reason behind the reduced attenuation performance as
the wheel speed is increased, the amplitude spectrum of the brake torque tra-
176 Chapter 6. BTV compensation with wheel speed measurements
140
← Compensation Started
120
Tb [Nm]
100
80
60
0 5 10 15 20 25 30
1400
Fcl [N]
1200
1000
800
0 5 10 15 20 25 30
Time [s]
(a) ωd = 1 Hz.
140
← Compensation Started
120
Tb [Nm]
100
80
60
0 2 4 6 8 10
1800
1600
Fcl [N]
1400
1200
0 2 4 6 8 10
Time [s]
(b) ωd = 3 Hz.
160
← Compensation Started
Tb [Nm] 140
120
100
80
0 1 2 3 4 5 6
1800
1600
1400
Fcl [N]
1200
1000
800
0 1 2 3 4 5 6
Time [s]
(c) ωd = 5 Hz.
160
← Compensation Started
140
120
Tb [Nm]
100
80
60
40
0 1 2 3 4 5 5.7
2000
1500
Fcl [N]
1000
500
0 1 2 3 4 5 5.7
Time [s]
(d) ωd = 7 Hz.
10
ωd = 1 Hz
5
0
0 1 2 3 4 5 6
Amplittude [Nm]
10
ωd = 3 Hz
5
0
0 2 4 6 8 10 12 14 16 18
20
ωd = 5 Hz
10
0
0 5 10 15 20 25 30
20
ωd = 7 Hz
10
0
0 5 10 15 20 25 30 35 40
Frequency [Hz]
jectories before and after attenuation are compared in Figure 6.10. The am-
plitude spectrum shows the amplitude distribution as a function of frequency,
and can be obtained using fast fourier transform (FFT). It is observed that
the amplitudes for the first harmonic (corresponds to the wheel speed) of the
brake torque responses decrease with compensation, indicating that the first-
order compensator is attenuating the first harmonic disturbance as intended.
No changes are expected from the second and third harmonics since the first-
order compensator only addresses the first harmonic of the BTV. However,
the attenuation action of the compensator also introduces low frequency noises
6.3. Results and discussions 179
150
← Compensation Started
Tb [Nm] 100
50
0 2 4 6 8 10 11.6
2000
Uncomp.
Fcl [N]
Comp.
1500
1000
0 2 4 6 8 10 11.6
6
ωd [Hz]
4
0 2 4 6 8 10 11.6
Time [s]
Figure 6.11: Experimentally obtained brake torque responses with and without
compensation under varying wheel speed.
with growing amplitudes as the wheel speed is increased, which leads to the
diminishing attenuation performance. This is due to the decreased emulation
performance of the forward-Euler discretisation method as the wheel speed
is increased. Nevertheless, the attenuated amplitudes outweighed the newly
introduced noise, as the RMSE of the compensated responses are decreased.
200
← Compensation Started
Tb [Nm]
100
0
0 5 10 15 20
2000
Fcl [N]
1000
Uncomp.
Comp.
0
0 5 10 15 20
5
ωd [Hz]
3
0 5 10 15 20
Time [s]
Figure 6.12: Experimentally obtained brake torque responses with and without
compensation under varying wheel speed and brake torque reference (thick
dash grey line).
6.4 Conclusions
A novel method for attenuating cold judder directly from the source of vi-
bration is proposed, where an electromechanical brake (EMB) is utilised to
generate a compensating clamp force to counteract the judder causing brake
torque variation (BTV). To this end, a BTV compensator that handles time-
varying disturbance frequencies is developed. Specific outcomes of this chapter
include:
linearised brake corner model incorporating the EMB and a disc rotor is
derived. The model assumes a bounded operation region and provides
the trace of brake torque given a clamp force trajectory.
• Utilising the brake corner model and the grey-box BTV model formulated
in Section 3.4, a BTV compensator scheduled using the measured wheel
speed and estimated acceleration is proposed. Output regulation theory
is employed to determine the existence of a compensator and to pro-
vide the feedforward gains for complete rectification of the BTV, where
a differential regulator equation is solved. Furthermore, the observer-
based controller structure is adapted for the proposed compensator by
incorporating the feedforward gains.
• A quadratic objective function is formulated to aid tuning the compen-
sator gains. Contrasting objectives of tracking speed and energy usage
are considered. Furthermore, the expected value of state estimation er-
rors are minimised. Tuning of the gains in the face of varying wheel
speed is handled using linear matrix inequality (LMI) formulations.
• To facilitate implementation of the compensator on a microcontroller,
discretisation using forward-Euler method and Tustin bilinear method
are investigated. Although the bilinear method provides better a emu-
lation performance, the computational load is overwhelming for current
implementation. On the other hand, the forward-Euler method is simple
to implement, albeit requires very small sampling period.
• The compensator is validated experimentally on the EMB and a dedi-
cated brake dynamometer, where the compensator is discretised using
forward-Euler method. Experimental results show that compensator is
able to reduce the root mean square error (RMSE) by up to 55%, and
the attenuation is consistent for both fixed and varying wheel speeds.
Better performance is expected if higher order compensators are tested.
However, this is limited by the stringent sampling rate requirement of
the forward-Euler method.
7
183
184 Chapter 7. BTV compensation with wheel position measurements
Brake corner model. The operating region of the LPV compensator stated
in Assumption 6.1 is adopted in this chapter for the adaptive feedforward
compensator design. Furthermore, by neglecting the noise in the brake corner
model (6.15), the nominal brake corner model is given by
x̄˙
=
A 0 x̄ Bu
+ ū,
ẇ 0 S(ωd ) w 0
(7.1)
h i x̄
eb = C Dew ,
w
7.1. Problem formulation 185
where the compensating control for the BTV is given by ū and the system
matrices A, Bu , C, Dew , S are given in (6.11) and (6.13). Note the overall clamp
force command is Fcl,r = ū + Fcl,f r (Tb,r ).
Problem 7.1 (BTV compensation with wheel position and speed measure-
ments). Given an operating region that satisfies Assumption 6.1, a fixed brake
torque reference Tb,r and a brake corner model (7.1), design a position, speed
and acceleration scheduled brake torque variation compensator (7.2) such that
the tracking error is attenuated, that is
lim eb (t) = 0,
t→∞
n o
for all x̄(0) ∈ R2 and w(0) ∈ w := [Td , w1T , . . . , wnT ]T : |Td | ≤ T d , Tbtv ∈ T ,
where Tbtv = [ 1 0 ]wk .
Pn
k=1
186 Chapter 7. BTV compensation with wheel position measurements
Remark 7.1. Problem 7.1 corresponds to the output regulation objective stated
in (b) of Problem 6.1. Although criteria for internal stability and optimal per-
formance are not specified in Problem 7.1 (which are analogous to (a) and (c)
in Problem 6.1), they will be addressed shortly in the following by Assump-
tion 7.1 and Proposition 7.1.
where χ := [eb , ėb ]T is the error state vector. The matrices B̄u and Γ(·) are
given by:
0
B̄u = , (7.5)
2µ̂rd kp
Γ(ωd , ω̇d ) = 1
2µ̂rd kp
[ −kp ωd2 −kp −(ω̇d +kd ωd ) ··· (nωd )2 −kp −n(ω̇d +kd ωd ) ]. (7.6)
7.2. Design of a BTV compensator with wheel position measurements 187
The A matrix can be made Hurwitz stable by proper selection of the near-
time-optimal clamp force controller (NTOC) gains. Therefore, the following
assumption can be made.
Assumption 7.1. The brake corner model (7.1) has a Hurwitz stable A ma-
trix.
Note Assumption 7.1 imposes internal stability on the tracking error system
(7.4). In other words, the tracking error system is stable in the absence of
BTV and compensating clamp force command, where w = 0 and ū = 0. To
enforce regulation of the tracking error eb in the presence of BTV, the adaptive
feedforward compensator (7.2) is adopted, where the compensating clamp force
command is given by
ū = Γ(ωd , ω̇d )ŵ,
The γ is estimated using the adaptive control law in (7.2). By defining the
following estimation error:
w̃ = ŵ − w, (7.8)
w̃˙ = ŵ˙ − ŵ
= Φ̇(θd )γ̂ + Φ(θd )γ̂˙ − Sw
= S(ωd )Φ(θd )γ̂ + gΦ(θd )$(θd )eb − Sw
= S(ωd )w̃ + gΦ(θd )$(θd )eb
= S(ωd )w̃ + gDew
T
eb , (7.10)
are used. By combining (7.9) and (7.10), the dynamics for both tracking error
and estimation error are given by
χ̇
=
A B̄u Γ(ωd , ω̇d ) χ
. (7.11)
w̃˙ gDew [ 1 0 ]
T
S(ωd ) w̃
Proposition 7.1. Consider the BTV Compensation Problem 7.1 and suppose
that Assumptions 6.1 and 7.1 hold. The adaptive feedforward compensator
(7.2) with a gain g ∈ R exponentially stabilises the error system (7.11) if there
exists a matrix P such that the following conditions are satisfied:
Ae,1 = Ae (ω d , αd ), Ae,2 = Ae (ω d , αd ),
Ae,3 = Ae (ω d , αd ), Ae,4 = Ae (ω d , αd ).
Furthermore, the lower bound of the decay rate in the error system (7.11)
satisfies the following optimisation problem:
max α (7.14a)
subject to P > 0, ATe,i P + P Ae,i + 2αP ≤ 0, i = 1, . . . , 4 (7.14b)
7.2. Design of a BTV compensator with wheel position measurements 189
Proof. The sufficient condition (7.13) follows directly from the definition of
quadratic stability (Boyd et al., 1994). Note the quadratic stability is a suffi-
cient (but not necessary) condition for the exponential stability of (7.11).
Consider a quadratic Lyapunov function V (ξ) = ξ T P ξ, where ξ =
[χT , w̃T ]T . Suppose dV (ξ)/dt ≤ −2αV (ξ) for all trajectories, then
V ξ(t) ≤ V ξ(0) e−2αt . (7.15)
into (7.15), where λmin (P ) and λmax (P ) denote that smallest and largest eigen-
values of P respectively, it can be shown that
By taking the square root of both sides of (7.16), it then follows that
r
−αt λmax (P )
kξ(t)k ≤ e λmin (P )
kξ(0)k
for all trajectories. Therefore the decay rate of the (7.11) is at least α. Note
the condition that dV (ξ)/dt ≤ −2αV (ξ) for all trajectories is equivalent to the
following LMI:
Procedure 7.1 (Bisection Algorithm). The lower bound of the decay rate, α
in (7.14) can be solved using the following steps:
Remark 7.2. Admittedly with some conservatism, the estimation of the lower
bound of the decay rate in the parameter-dependent error system (7.11) can
be reduced to a LMI problem and hence is numerical tractable. The estimated
value can then be employed for tuning of the observer gain, g.
Note the compensator structure (7.2) can be generalised by consid-
ering a compensator gain matrix with diagonal elements, where g :=
diag (g1 , . . . , g2n+1 ) ∈ R(2n+1)×(2n+1) . The additional flexibility allows fine tun-
ing of the compensation performance.
0.09
0.08
0.07
Lower Bound of Decay Rate, α
0.06
0.05
0.04
0.03
0.02
0.01
0
0 1 2 3 4 5 6 7 8 9 10
Compensator Gain, g
Figure 7.2: The lower bound of the decay rate, α versus the compensator gain,
g.
error decay rate is plotted against compensator gains in the range of [0, 10]
in Figure 7.2. It is shown that the compensator gain g = 5 yields the largest
decay rate.
Furthermore, an experimentally obtained BTV profile illustrated in Fig-
ure 6.2 is employed to simulate the BTV attenuation performance of the pro-
posed compensator. Figure 7.3 illustrates the BTV compensation for first-
order compensators with different gains, where g = {1, 5, 10}. A fixed brake
torque reference, Tb,r = 100 Nm is given. The compensator with g = 5 reaches
steady state in the shortest time as expected. However, it also yields a larger
overshoot compared to the compensator with g = 1. The compensator with
g = 10 has the largest overshoot. Furthermore, the conservativeness of Fig-
ure 7.2 can be seen here, where the α for g = 10 is zero, hence suggesting that
the response may be unstable. However, the simulated trajectory in Figure 7.3
shows that the response is stable.
Figure 7.4 demonstrates the BTV attenuation responses for compensators
with a compensator gain g = 3 and different orders of harmonics. As ex-
192 Chapter 7. BTV compensation with wheel position measurements
200
150
Tm [Nm]
100
50
0
0 0.5 1 1.5 2 2.5 3 3.5 4
1500
1000
Fcl [N]
500
pected, the third-order compensator achieves the best BTV attenuation at the
expense of higher computational load. Compared to the first-order compen-
sator, a significant improvement in attenuation is observed in the second-order
compensator. However, compared to the second-order compensator, the im-
provement achieved by the third-order compensator is insignificant and does
not justify the increased in implementation complexity.
200
150
Tm [Nm]
100
50
0 0.5 1 1.5 2 2.5 3 3.5 4
1500
1000
Fcl [N]
500
tive feedforward BTV compensator can operate at a lower sampling rate com-
pared to the LPV compensator. Additionally, due to the lower sampling rate
requirement, a higher order discretised adaptive feedforward compensator can
be implemented, therefore achieving a better BTV attenuation.
120 120
100 100
80 80
0 1 2 3 4 0 0.5 1 1.5
140 140
Tb [Nm]
120 120
100 100
80 80
0 1 2 3 4 0 0.5 1 1.5
Time [s] Time [s]
waveform revealed that the first harmonic has the largest amplitude, followed
by smaller amplitudes of the second and the third harmonics. This suggests
that BTV compensators of up to three orders are sufficient in implementation
for this case. First, experiment results with fixed wheel speed are presented.
This scenario frequently occurs when light braking is applied while a vehicle is
descending a ramp. This is then followed by experiment results with general
braking profiles.
Fixed wheel speed. Figure 7.6 illustrates the BTV compensation for two
different wheel speeds, where ωd = {1, 7} Hz. The test sequence had a fixed
brake torque reference at 150 Nm and began without BTV compensation. The
first-order compensator was started at 4 seconds, followed by the second-order
compensator at 8 seconds, and lastly the third-order compensator was started
at 12 seconds. Due to the variability of friction coefficient, the DC offset of the
uncompensated brake torque trajectories differ from the setpoint at 150 Nm.
As soon as the first-order compensator was switched on, the BTV amplitude
decreased. Additionally, the DC offset reduced and the trajectory was centred
7.3. Results and discussions 195
at the setpoint thanks to the integral action within the compensator. Further-
more, additional attenuation was observed when the second-order compensator
was switched on. However, the improvement in BTV attenuation of the third-
order compensator was visually insignificant compared to the second-order
compensator.
To qualitatively assess the BTV attenuation, the root mean square error
(RMSE) of the brake torque tracking trajectories are calculated and listed in
Table 7.1. To remove the DC offsets and isolate the sinusoidal variations in
brake torque trajectories, the RMSEs without the DC offsets are also calcu-
lated. Additionally, a dimensionless quantity can be obtained by normalising
the RMSEs (without DC bias) of the compensated responses with respect to
the uncompensated trajectories, where the Normalised RMSE is given by
RMSE w/o DC Offset of Compensated Response
Normalised RMSE = .
RMSE w/o DC Offset of Uncompensated Response
For the 1 Hz wheel speed scenario, the first-order compensator reduced the
normalised RMSE by 37%, and the second-order compensator achieved another
26% improvement. However, the improvement obtained by the third-order
compensator was only 4%, and is insignificant from a practical perspective.
Similar trend was also observed for the 7 Hz wheel speed scenario, where the
first-order compensator reduced the normalised RMSE by 34%, followed by
another 31% for the second-order compensator, and a 1% improvement for the
third-order compensator.
Compared to the LPV compensator operating at 7 Hz wheel speed (Fig-
ure 6.9), the BTV attenuation attained by adaptive feedforward compensator
196 Chapter 7. BTV compensation with wheel position measurements
250
← Uncomp. → ← 1st order → ← 2nd order → ← 3rd order →
200
Tb [Nm]
150
100
0 2 4 6 8 10 12 14 16
1000
900
800
Fcl [N]
700
600
500
0 2 4 6 8 10 12 14 16
Time [s]
(a) ωd = 1 Hz.
250
← Uncomp. → ← 1st order → ← 2nd order → ← 3rd order →
200
Tb [Nm]
150
100
0 2 4 6 8 10 12 14 16
1500
Fcl [N]
1000
500
0 2 4 6 8 10 12 14 16
Time [s]
(b) ωd = 7 Hz.
is significantly better visually (even though a lower brake torque setpoint was
used in the LPV compensator test case, where the BTV was smaller). Note
a second-order adaptive feedforward compensator can be practically imple-
mented. However, the second-order LPV compensator cannot be practically
implemented in the current setup due to the high sampling rate requirement.
Amplitude spectrum of the uncompensated and compensated responses are
depicted in Figure 7.7. The amplitude spectrum is used to illustrate the ampli-
tude distribution of a signal as a function of frequency. As the BTV is caused
by the DTV, the first harmonic of the uncompensated brake torque response
has the largest amplitude, followed by the second and third harmonics. As
expected, when the first-order compensator is activated, the amplitude of the
first harmonic (of the compensated brake torque response) is reduced. How-
ever, the amplitude of the second harmonic remains unchanged. When the
second-order compensator is activated, the peaks for both the first and the
second harmonics are attenuated (the peak for the second harmonic is below
1 Nm, and is omitted from the plot). Note the first-order and second-order
compensators cannot attenuate amplitudes for the third and higher order har-
monics.
Variable wheel speed. Figure 7.8 shows the brake torque compensation
under a varying wheel speed, starting from 6 Hz while simulating a light brak-
ing event with 2 rad/s2 deceleration. It is demonstrated that the second-order
compensator achieves 44% reduction in normalised RMSE even under varying
speed. Furthermore, the DC bias found in the uncompensated response was
also corrected by the compensator.
7.4 Conclusions
An adaptive feedforward brake torque variation (BTV) compensator is pro-
posed in this chapter, addressing the need of high sampling rate required by
the linear parameter-varying (LPV) BTV compensator developed in Chap-
ter 6. The reduced sampling rate requirement of the adaptive feedforward
compensator allows BTV compensation at higher wheel speed and attenua-
tion of BTV with higher order harmonics. The main contributions of this
chapter are summarised as follows:
198 Chapter 7. BTV compensation with wheel position measurements
15
ωd = 1 Hz
Amplitude [Nm]
10
0
0 1 2 3 4 5 6
Frequency [Hz]
15
ωd = 7 Hz
Amplitude [Nm]
10
0
0 5 10 15 20 25 30 35 40
Frequency [Hz]
Figure 7.7: Amplitude spectrum of brake torque trajectories for different wheel
speeds. The peaks for uncompensated responses are represented by circles,
peaks for compensated responses using the first-order and the second-order
compensators are represented by squares and diamonds respectively.
300
← Compensation Started
Tb [Nm]
200
100
0 2 4 6 8 10 12
2000
Fcl [N]
1000
Uncomp. 2nd order
0
0 2 4 6 8 10 12
6
ωd [Hz]
5
4
3
0 2 4 6 8 10 12
Time [s]
Figure 7.8: Experimentally obtained brake torque responses with and without
compensation under varying wheel speed.
8
Conclusions
This thesis investigates control techniques for clamp force tracking and brake
torque variation compensation for a production-ready prototype automotive
electromechanical brake (EMB). The review of existing control approaches in
Chapter 2 identified the gap in literature for the active compensation of brake
torque variation, which is the main cause of brake judder. To support control
development for the EMB, a rapid parameter identification for a reduced-
order EMB model was proposed. This was followed by the development of
high performance clamp force controllers, where the time-optimal control, ro-
bust near-time-optimal control (NTOC) and explicit model predictive con-
trol (MPC) were considered. The NTOC was implemented on a production
ready prototype EMB and showed favourable performance compared to the
industry-tuned cascaded proportional integral controller. The achieved high
closed-loop bandwidth of this design was then employed as the inner control
loop in the brake torque compensator design. Two compensation methods, the
linear parameter-varying (LPV) and the adaptive feedforward compensators
were examined. The three specific research aims stated in Section 2.5 were an-
swered throughout Chapters 4 to 7. In the following, the main contributions
are summarised and future research directions are identified.
201
202 Chapter 8. Conclusions
8.1 Contributions
Developed a rapid EMB parameter identification procedure
A rapid parameter identification procedure for the EMB was proposed, which is
intended to support quick model-based controller tunings and post-production
quality control checks. The optimal experimental design problem was formu-
lated and an optimal experimental trajectory was determined with the aim of
minimising the experimental time and maximising parameter estimation ac-
curacy. Additionally, two parameter estimation techniques, the output error
method and prediction error method were investigated in order to infer the
model parameters from experimental measurements. It was observed that the
prediction error method provides a marginally better estimation accuracy com-
pared to the output error method but requires significantly longer evaluation
time. Compared to the incumbent approach, the proposed rapid parameter
identification procedure significantly reduces the time required for experiments
from nine hours to one minute.
rise time of the NTOC was shown to be typically within 5% of the time-optimal
control, indicating that the NTOC is able to almost fully utilise the available
motor capacity to achieve a rapid setpoint tracking without causing control
chatter. Furthermore, the NTOC better utilises the available motor torque
compared to the PI control, and is significantly faster for small setpoints.
Additionally, a holistic control design approach incorporating tracking er-
ror, control effort and system constraints was examined under the model
prediction control (MPC) framework. Although this design approach shows
greater flexibility and slight control performance improvements compared to
the NTOC, the intensive computational and large memory demands are pro-
hibitive under current hardware limits.
• Develop and extend the output regulation theory for a time-varying dis-
crete system. This result will be beneficial to the development of a
discrete LPV BTV compensator, and possibly reducing the stringent
sampling rate requirement needed for implementation.
• While the adaptive feedforward scheme has been successfully imple-
mented on the EMB system to attenuate the BTV, which can be per-
ceived as an output disturbance, it is unclear whether the current ap-
proach can be generalised to any output regulation problem. A general-
isation of this design approach will be useful for solving a wide range of
vibration attenuation problems.
• In the current work, the proposed BTV compensators have been exper-
imentally validated on the dedicated brake dynamometer. However, it
will be interesting to validate this on a larger brake dynamometer with
higher speed and larger torque capacity. Furthermore, disc rotors with
different DTV waveforms shall be used in the elaborated tests. A stan-
dardised testing procedure should be used, for example, the SAE J3002
“Dynamometer Low-Frequency Brake Noise Test Procedure” (which is
currently a standard work in progress and has been not released to the
public).
• It is noted that the proposed BTV compensator designs utilise brake
torque and/or disc angular position measurements from the dynamome-
ter, which may be lacking in actual implementation in a production vehi-
cle. While it is not uncommon to have brake torque measurement setup
in prototype vehicles, equipping this sensor in a near future production
vehicle is not expected. Nevertheless, brake torque may be estimated us-
ing the available vehicle acceleration measurements. Furthermore, disc
position may be approximated by integrating the measured wheel speed.
• The concept of utilising a brake to actively attenuate brake judder maybe
extended to hydraulic brakes, and the feasibility of this extension war-
rants further investigations.
• Braking action is abrasive in nature. Therefore, it remains a distinct
possibility that the disc thickness variation causing the judder can be
machined away using the brake calliper itself. The research questions
are what type of braking scenarios this approach might be possible, and
the required extensions to the original compensation schemes.
A
Periodic signals
.
A periodic signal d(·) cyclical in domain X with period p can be charac-
terised as
d(x + p) = d(x), ∀x ∈ X , x + p ∈ X .
It can be decomposed into the sum of a set of simple oscillating functions using
Fourier series (Kreyszig, 2006), which carries the form of
∞
d(x) = a0 + ak cos (kωx) + bk sin (kωx) ,
X
k=1
a0 = d(x) dx,
p − p2
2Z 2
p
k=−∞
205
206 Appendix A. Periodic signals
Periodic signal
Initial function d(t)
−ps
+ e
t t
−p 0 + 0 p 2p 3p
dq d(t) q−1
X dk d(t)
+ γk = 0. (A.3)
dtq k=0 dtk
k=0
For a periodic disturbance, the generating polynomial (A.4) will carry the form
of the denominator in (A.2).
where
a = A cos (φ), b = −A sin (φ).
If the following new variables are defined:
A cos (ωt + φ)
xd1 = ,
−Aω sin (ωt + φ)
then the following model for the sinusoidal disturbance (A.5) can be obtained:
0 1
= 2 xd1 ,
ẋ
d1
Σ1 : −ω 0 (A.6)
h i
d = 1 0 xd1 .
h̄ = αh h,
where
10000−1 0
αh = .
0 100−1
The Jacobian ∂fd
∂xT
is given by:
d
∂fd h i
= ∂fd ∂fd
,
∂xTd ∂θm ∂ωm
where
∂fd 1
= ts ∂Fcl ,
∂θm − J ∂θm N + µs tanh (εωm )
1+(ωm /ωs )2
" ts #
∂fd
=
.
∂ωm − tJs Tv +ε TC +
Ts −TC
1+(ωm /ωs )2
(1−tanh (εωm )2 )− 2ωmω(T2s(−TC ) tanh (εωm )
1+ω 2 /ω 2 )
2 +1
s m s
209
210 Appendix B. Expansion of equation 4.11
Note that the differentiation of the stiffness with respect to the motor position
is given by:
∂Fcl
= 3k1 θm2
+ 2k2 θm .
∂θm
Furthermore, the stiction is governed by:
Ts = Ts0 + µs Fcl .
The Jacobian ∂ h̄
∂xT
is given by:
d
∂ h̄ ∂Fcl
0
= α h
∂θm .
∂xdT
0 1
The expressions for ∂fd
∂ θ̄i
are procured as:
∂fd 0
= 3 tanh (εω )
µs k1,nom θm
∂ θ̄1 − tJs N k1,nom θm
3
+ 1+ωm 2 /ω 2
m
s
∂fd 0
= 2 tanh (εω )
µs k2,nom θm
∂ θ̄2 − tJs N k2,nom θm
2
+ 1+ωm 2 /ω 2
m
s
∂fd 0
=
∂ θ̄3 − tJs (TC,nom tanh (εωm )) 1 − 2 /ω 2
1+ωm
1
s
∂fd 0
=
∂ θ̄4 − tJs (Ts0,nom tanh (εωm )) 1+ω21 /ω2
m s
∂fd 0
=
∂ θ̄5 − tJs Tv,nom ωm
∂fd 0
=
∂ θ̄6 − tJs (µs,nom tanh (εωm )) 1+ωF2cl/ω2
m s
∂fd 0
= ts .
∂ θ̄7 − J (2ωs,nom ωm tanh (εωm )) ω3 (1+ω2 /ω2 )2
2 Ts −TC
s m s
Baumann, D., Hofmann, D., Vollert, H., Nagel, W., Henke, A., Foitzik, B. and
Goetzelmann, B. (2011), ‘Self boosting electromechanical friction brake’,
United States Patent US8002088B2.
211
212 Bibliography
Doi, K., Mibe, T., Matsui, H., Tamasho, T. and Nakanishi, H. (2000), ‘Brake
judder reduction technology–brake design technique including friction ma-
terial formulation’, JSAE review 21(4), pp. 497–502.
Duan, C. and Singh, R. (2011), ‘Analysis of the vehicle brake judder problem
by employing a simplified source–path–receiver model’, Proceedings of the
Institution of Mechanical Engineers, Part D: Journal of Automobile Engi-
neering 225(2), pp. 141–149.
Haigh, M. J., Smales, H. and Abe, M. (1993a), Vehicle judder under dynamic
braking caused by disc thickness variation, in ‘Braking of Road Vehicles’,
pp. 247–258. IMechE paper C444/022/93.
Haigh, M., Smales, B. and Abe, M. (1993b), RTV - A friction material designers
view, in ‘SAE Brake Colloquium and Engineering Display’. SAE Paper No.
933070.
Hara, S., Yamamoto, Y., Omata, T. and Nakano, M. (1988), ‘Repetitive con-
trol system: a new type servo system for periodic exogenous signals’, IEEE
Transactions on Automatic Control 33(7), pp. 659–668.
Inoue, T., Nakano, M. and Iwai, S. (1981), High accuracy control of servomech-
anism for repeated contouring, in ‘Annual Symposium on Incremental Mo-
tion Control Systems and Devices’, pp. 258–292.
Bibliography 215
Kao, T. K., Richmond, J. W. and Douarre, A. (2000), ‘Brake disc hot spotting
and thermal judder: an experimental and finite element study’, International
Journal of Vehicle Design 23(3/4), pp. 276–296.
Kim, M.-G., Jeong, H.-I. and Yoo, W.-S. (1996), Sensitivity analysis of chassis
system to improve shimmy and brake judder vibration on steering wheel, in
‘SAE International Congress and Exposition’. SAE Paper No. 960734.
216 Bibliography
Kubota, M., Suenaga, T. and Doi, K. (1998), A study of the mechanism causing
high-speed brake judder, in ‘SAE International Congress and Exposition’.
SAE Paper No. 980594.
Lee, K. and Brooks, Jr, F. W. (2003), ‘Hot spotting and judder phenomena in
aluminium drum brakes’, Journal of Tribology 125, pp. 44–51.
Bibliography 217
Makkar, C., Dixon, W., Sawyer, W. and Hu, G. (2005), A new continuously dif-
ferentiable friction model for control systems design, in ‘2005 IEEE/ASME
International Conference on Advanced Intelligent Mechatronics’, pp. 600–
605.
McGill, R. (1965), ‘Optimum control, inequality state constraints, and the gen-
eralized newton-raphson algorithm’, SIAM Journal of Control 3(2), pp. 291–
298.
Nissan North America (2005), ‘Warranty extension for “brake judder” condi-
tion involving model year 2005 and 2005 Titan and Armada vehicles’, Claims
Bulletin. Reference: WB/05-011.
Bibliography 219
Olsson, H., Åström, K. J., Canudas de Wit, C., Gäfvert, M. and Lischinsky,
P. (1998), ‘Friction models and friction compensation’, European Journal of
Control 4(3), pp. 176–195.
Pao, L. Y. (1994), Characteristics of the time-optimal control of flexible struc-
tures with damping, in ‘Proceedings of the IEEE Conference on Control
Applications’, pp. 1299–1304.
Pao, L. Y. and Singhose, W. E. (1998), ‘Robust minimum time control of
flexible structures’, Automatica 34(2), pp. 229–236.
Pontryagin, L. S., Boltyanskii, V. G., Gamkrelidze, R. V. and Mishchenko,
E. F. (1962), The Mathematical Theory of Optimal Processes, John Wiley
and Sons.
Raol, J. R., Girija, G. and Singh, J. (2004), Modelling and parameter estima-
tion of dynamic systems, Institution of Engineering and Technology.
Reuter, D. F., Lloyd, E. W., Zehnder, J. W. and Elliott, J. A. (2003), Hy-
draulic design considerations for EHB systems, in ‘SAE World Congress
and Exhibition’, Detroit, MI. SAE Paper No. 2003-01-0324.
Roberts, R., Gombert, B., Hartmann, H. and Schautt, M. (2003), Modelling
and validation of the mechatronic wedge brake, in ‘SAE Brake Colloquium
and Exhibition’. SAE Paper No. 2003-01-3331.
Roozen, N. B., Koevoets, A. H. and den Hamer, A. J. (2008), ‘Active vi-
bration control of gradient coils to reduce acoustic noise of MRI systems’,
IEEE/ASME Transactions on Mechatronics 13(3), pp. 325–334.
Rosen, J. B. (1966), ‘Iterative solution of nonlinear optimal control problems’,
SIAM Journal on Control 4(1), pp. 223–244.
Ryan, E. P. (1980), ‘On the sensitivity of a time-optimal switching function’,
IEEE Transactions on Automatic Control 25(2), pp. 275–277.
Saberi, A., Stoorvogel, A. and Sannuti, P. (2000), Control of linear systems
with regulation and input constraints, Springer.
Sacks, A., Bodson, M. and Khosla, P. (1996), ‘Experimental results of adaptive
periodic disturbance cancellation in a high performance magnetic disk drive’,
Journal of Dynamic Systems, Measurement, and Control 118(3), pp. 416–
424.
SAE-Australasia (2007), ‘Special commendation’, Autoengineer 29, pp. 19.
Salton, A. T., Chen, Z. and Fu, M. (2012), ‘Improved control design methods
for proximate time-optimal servomechanisms’, IEEE/ASME Transactions
on Mechatronics 17(6), pp. 1049–1058.
220 Bibliography
Schwarz, R., Isermann, R., Böhm, J., Nell, J. and Rieth, P. (1998), Model-
ing and control of an electromechanical disk brake, in ‘SAE International
Congress and Exposition’, Detroit, MI. SAE Paper No. 980600.
Schwarz, R., Isermann, R., Böhm, J., Nell, J. and Rieth, P. (1999), Clamp-
ing force estimation for a brake-by-wire actuator, in ‘SAE International
Congress and Exposition’. SAE Paper No. 1999-01-0482.
Shim, H., Kim, J.-S., Kim, H. and Back, J. (2010), ‘A note on the differential
regulator equation for non-minimum phase linear systems with time-varying
exosystems’, Automatica 46(3), pp. 605–609.
Shim, H., Lee, J., Kim, J.-S. and Back, J. (2006), Output regulation problem
and solution for LTV minimum phase systems with time-varying exosystem,
in ‘SICE-ICASE International Joint Conference’, pp. 1823–1827.
Sri-Jayantha, S. M., Dang, H., Sharma, A., Yoneda, I., Kitazaki, N. and Ya-
mamoto, S. (2001), ‘Truetrack™ servo technology for high tpi disk drives’,
IEEE Transactions on Magnetics 37(2), pp. 871–876.
Stringham, W., Jank, P., Pfeifer, J. and Wang, A. (1993), Brake roughness-
disc brake torque variation, rotor distortion and vehicle response, in ‘SAE
International Congress and Exposition’. SAE Paper No. 930803.
Tamasho, T., Doi, K., Hamabe, T., Koshimizu, N. and Suzuki, S. (2000),
‘Technique for reducing brake drag torque in the non-braking mode’, JSAE
review 21(1), pp. 67–72.
von Groll, M., Mueller, S., Meister, T. and Tracht, R. (2006), ‘Disturbance
compensation with a torque controllable steering system’, Vehicle System
Dynamics 44(4), pp. 327–338.
Wang, N., Kaganov, A., Code, S. and Knudtzen, A. (2005), ‘Actuating mech-
anism and brake assembly’, International Patent No. WO 2005/124180 A1.
Wang, Y., Gao, F. and Doyle, F. J. (2009), ‘Survey on iterative learning con-
trol, repetitive control, and run-to-run control’, Journal of Process Control
19(10), pp. 1589–1600.
Whitcomb, R. (2004), Avro Aircraft and Cold War Avaition, Vanwell Publish-
ing.
Zhang, L., Ning, G. and Yu, Z. (2007), Brake judder induced steering wheel
vibration: Experiment, simulation and analysis, in ‘SAE Brake Colloquium
and Exhibition’. SAE Paper No. 2007-01-3966.
Zhang, Y. (2000), Stability and performance tradeoff with discrete time tri-
angular search minimum seeking, in ‘Proceedings of the American Control
Conference’, vol. 1, pp. 423–427.
Zhang, Z. and Serrani, A. (2006), ‘The linear periodic output regulation prob-
lem’, Systems & control letters 55(7), pp. 518–529.
Zheng, J., Guo, G., Wang, Y. and Wong, W. E. (2006), ‘Optimal narrow-band
disturbance filter for pzt-actuated head positioning control on a spinstand’,
IEEE Transactions on Magnetics 42(11), pp. 3745–3751.
Zhou, J., Zhou, R., Wang, Y. and Guo, G. (2001), ‘Improved proximate time-
optimal sliding-mode control of hard disk drives’, IEE Proceedings Control
Theory and Applications 148(6), pp. 516–522.
Author/s:
Lee, Chih Feng
Title:
Brake force control and judder compensation of an automotive electromechanical brake
Date:
2013
Citation:
Lee, C. F. (2013). Brake force control and judder compensation of an automotive
electromechanical brake. PhD thesis, Department of Mechanical Engineering, The
University of Melbourne.
Publication Status:
Unpublished
Persistent Link:
http://hdl.handle.net/11343/38499